Sie sind auf Seite 1von 55

Accepted Manuscript

Recent Progress in Organic Redox Flow Batteries: Active Materials,


Electrolytes and Membranes http://www.journals.elsevier.com/
journal-of-energy-chemistry/
Hongning Chen , Guangtao Cong , Yi-Chun Lu

PII: S2095-4956(17)31105-1
DOI: 10.1016/j.jechem.2018.02.009
Reference: JECHEM 544

To appear in: Journal of Energy Chemistry

Received date: 8 December 2017


Revised date: 1 February 2018
Accepted date: 4 February 2018

Please cite this article as: Hongning Chen , Guangtao Cong , Yi-Chun Lu , Recent Progress in Or-
ganic Redox Flow Batteries: Active Materials, Electrolytes and Membranes, Journal of Energy Chem-
istry (2018), doi: 10.1016/j.jechem.2018.02.009

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service
to our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and
all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Recent Progress in Organic Redox Flow Batteries: Active


Materials, Electrolytes and Membranes
Hongning Chen,# Guangtao Cong# and Yi-Chun Lu*
Electrochemical Energy and Interfaces Laboratory, Department of Mechanical and Automation
Engineering, The Chinese University of Hong Kong, Shatin, N.T. 999077, Hong Kong SAR,
China.

T
*
Corresponding Author: Yi-Chun Lu. E-mail: yichunlu@mae.cuhk.edu.hk

IP
#
These authors contribute equally

CR
Abstract

US
Redox flow batteries (RFBs) have great potentials in the future applications of both large scale
energy storage and powering the electrical vehicle. Critical challenges including low volumetric
AN
energy density, high cost and maintenance greatly impede the wide application of conventional
RFBs based on inorganic materials. Redox-active organic molecules have shown promising
prospect in the application of RFBs, benefited from their low cost, vast abundance, and high
M

tunability of both potential and solubility. In this review, we discuss the advantages of redox
active organic materials over their inorganic compart and the recent progress of organic based
ED

aqueous and non-aqueous RFBs. Design considerations in active materials, choice of electrolytes
and membrane selection in both aqueous and non-aqueous RFBs are discussed. Finally, we
PT

discuss remaining critical challenges and suggest future directions for improving organic based
RFBs.
CE

Keywords: Organic materials; electrolyte; membrane; flow battery; low cost.


AC

1
ACCEPTED MANUSCRIPT

1. Introduction
Energy storage is an indispensable bridge between intermittent renewable power sources and
energy demands. The widespread and deep penetration of renewable energy relies on low-cost
and efficient energy storage technologies[1, 2]. The United Nations (UN) announced a record-
breaking 138.5 GW of renewable energy added worldwide in 2016, which represents an

T
investment of $242 million, especially for solar energy and wind energy[3]. However, the

IP
renewable energies such as solar energy and wind energy are highly intermittent, which requires

CR
reliable energy storage technologies to store the intermittent energy at peak time and compensate
the demand at valley time[4-6]. Redox flow batteries (RFBs) are one of the most promising
energy storage technologies for large-scale applications[7-9]. A typical RFB consists of three

US
main components including cell stack, external tanks and flow pipes, as shown in Figure 1. With
the electrolytes stored in the external tanks, one of the most attractive advantages of the RFBs is
AN
the flexibility in decoupling of power and energy[10-14]. In a RFB system, the amount of stored
energy scales with the volume of the tank while the power rating scales with the size of cell
stack[15-19].
M

Vanadium redox flow battery (VRFB) is one of the most successful RFB technologies. In a
ED

typical VRFB, VO2+/VO2+ and V2+/V3+ are separated as catholyte active species and anolyte
species, respectively, and the reactions are shown as follows[10]:
PT

Catholyte reaction:

Anolyte reaction:
CE

Cell reaction:
AC

2
ACCEPTED MANUSCRIPT

T
IP
CR
US
Figure 1. A schematic illustration of redox flow batteries.
AN
VRFB was first reported by Skyllas-Kazacos and her co-workers at the University of New South
Wales in the 1980s, and later was widely developed worldwide owing to its unique advantage in
M

alleviating crossover since only one type of active element (vanadium) was used in this cell[20-
22]. Recently a long-life VRFB of more than 200,000 cycles has been demonstrated in Japan for
ED

over 10 years[23]. However, the energy density of VRFB is considerably low (20-60 Wh/L [24,
25]) due to the low solubility of vanadium ion in the electrolyte (2.0 – 3.0 M)[7, 10, 26] and the
price of vanadium is high ($27/kg, Chemicool. http://www.chemicool.com/, Accessed August 21,
PT

2017), which strongly hinder the application of VRFB in many fields. In addition to vanadium,
other inorganic active species have been applied in RFBs [10, 27-29].
CE
AC

3
ACCEPTED MANUSCRIPT

2. The focus of the review


The application of organic materials in RFBs has been identified one of the most important
approaches to address the critical challenges facing inorganic RFBs. The advantages of applying
organic materials in RFBs lie in its diverse molecular moieties, high structural tunability, and
low material cost. Three recent review articles[30-32] have summarized the development in
organic redox flow batteries with a major focus on redox active materials. In this review, we will

T
first discuss the advantages of organic materials for RFBs compared with inorganic-based RFBs;

IP
second, we will discuss the recent progress of organic RFBs in redox active materials, the
properties of the electrolyte and the design of the membrane including polymeric and ceramic

CR
membranes. Different design considerations in active materials, electrolyte and membrane
designs in aqueous and non-aqueous electrolyte media will be discussed. Finally, we will discuss

US
remaining critical challenges and suggest future directions for improving organic based RFBs.

3. Advantages of organic materials in RFBs


AN
With the rapid growing demand of renewable energy and prosperity of new electrical vehicle
market, the need for low-cost, earth-abundant, high-energy and long-lasting energy storage
M

devices is vast[33-36]. Existing inorganic-based RFBs are suffering from high-cost and low
energy density. Applying organic materials in RFBs promises several advantages[10, 11, 37]:
ED

(1). Diverse molecular moieties. Organic materials offer a large variety of molecules as active
materials in RFBs[31, 38, 39]. For example, the metallocenes (e.g. ferrocene (Fc), chromocene
PT

etc.) which consist of metal ion and aromatic cyclopentadienyl anions (Cp, C5H5-) are one of the
most commonly used organic active materials in RFBs owing to the excellent electrochemical
CE

reversibility and stability[40-42]. In addition, nitroxide radicals (e.g. 2,2,6,6-


tetramethylpiperidin-1-yl)oxyl (TEMPO)[15, 43], 2-phenyl-4,4,5,5-tetramethylimidazoline-1-
oxyl-3-oxide (PTIO) etc. [44, 45]), quinone-based materials (e.g. 9,10-anthraquinone-2,7-
AC

disulphonic acid (AQDS)[46], (1,2-benzoquinone-3,5-disulfonic acid) BQDS[47] etc.),


heterocyclic aromatics (e.g. quinoxalines[48, 49], 2,1,3-benzothiadiazole (BzNSN)[50], etc.),
carbonyl compounds (e.g. ( 9-fluorenone) FL[51], 2,5-di-tert-butyl-1,4-bis(2-
methoxyethoxy)benzene (DBBB)[52] etc.) and many other organic molecules have been
demonstrated redox active and applied in organic RFBs.

4
ACCEPTED MANUSCRIPT

(2). High structure tunability and wide range of redox potentials. The diverse selection of
organic molecules provides wide range of reaction potentials from -0.5 V vs Li/Li+ to 5 V vs
Li/Li+. In addition to the redox active center, the high structure tunability of the organic materials
provides facile methods to tune the redox potential of the molecule via functional group[30-32].
In general, the redox potential can be elevated with the addition of electron-withdrawing groups
(e.g. COOH, =O, OH)[53] and lowered with the addition of electron-donating groups (e.g. R, OR,

T
NR2, in which R is an alkyl)[53]. In search for candidates at the high potential side, organic

IP
materials which are exploited in Li-ion battery as overcharge protection typically have a high
reaction potential up to 4.8 V vs Li/Li+. For instance, tetraethyl-2,5-di-tert-butyl-1,4-phenylene

CR
diphosphate (TEDBPDP) has a reversible reaction potential around 4.80 V vs Li/Li+ [54]. At the
low potential side, K. Gong et al., reported that biphenyl (BP) molecule can be reversibly

US
reduced to form the biphenyl radical anion (BP-) at a very negative reaction potential of -2.70 V
vs. SCE (or -2.46 V vs. SHE) in acetonitrile. Consequently, a much higher cell voltage can be
achieved using organic materials compared to inorganic based RFBs[17]. The cell voltage is
AN
critical in improving the energy density of RFBs and increase the competitiveness of RFBs for
both large-scale grid storage and automobile applications.
M

(3). Material abundance and low material costs. Many organic active materials can be
obtained from the abundant natural resources such as food and plants[55, 56]. The abundant
ED

resources can dramatically decrease the costs of the RFB system. For example, the V2O5 which
is the raw material for VRFB has a price of 10-12 $/kg[57]. However, quinone-based active
PT

materials which are well-known organic molecules in bio-logical electron-transfer processes cost
6 $/kg[30, 47]. Additionally, K. Lin et al. introduced a family of organic molecules for RFB
CE

applications which can be derived from vitamin B2[58]. In addition to low-cost, organic
materials have advantages in abundant reserves, safety and stability[59-61].
AC

4. Aqueous based organic RFBs


Aqueous based RFBs provide fast kinetics and high ionic conductivity compared with non-
aqueous based RFBs[62-64]. For large-scale energy storage applications, the relatively low cost
of aqueous electrolyte compared to non-aqueous electrolyte adds additional competitiveness.
However, the active materials in conventional aqueous RFBs are limited in choice and some of
them are expensive, which significantly increase the total cost of the system. The development of

5
ACCEPTED MANUSCRIPT

aqueous based organic RFBs offers a new and promising direction for the future low-cost energy
storage technique[32, 46]. Here the main progress of aqueous based organic RFBs is reviewed
including active materials, electrolytes and membranes.

4.1. Active materials

There are mainly two types of organic active materials used in RFBs including metal-containing

T
organic active materials and metal-free organic active materials. The charge-transfer reactions of

IP
metal-containing active materials are mainly based on the metal ions in the molecule, which
typically exhibit fast reaction kinetics. However, the redox potential of metal-containing organic

CR
active materials is similar to their metal centers, which limits the potential range of the metal-
containing organic materials. On the other hand, the metal-free organic active materials have

US
much broader potential windows. In addition, the metal-free organic active materials can provide
multiple electron transfers, which improves the energy density. However, the solubility of many
AN
existing metal-free organic active materials is still low, which seriously limits their energy
density.
M

4.1.1. Metal-containing organic active materials.


ED

Metallocenes are the most commonly used metal-containing organic active materials in aqueous
RFBs, which are sandwich compounds consisting of two aromatic cyclopentadienyl anions (Cp,
PT

C5H5-) bound to a metal ion center. Metallocenes have high stability and reversibility in both
aqueous and non-aqueous systems[65]. However, the solubility of metallocenes in aqueous
CE

solvents is low[13, 41]. For example, ferrocene derivatives are barely soluble in water (< 100
mM)[40]. In 2016, B. Hu et al. firstly introduced a highly water-soluble ferrocene molecule (4 M
(ferrocenylmethyl)trimethylammonium chloride (FcNCl) and 3.1 M N1-ferrocenylmethyl-
AC

N1,N1,N2,N2,N2-pentamethylpropane-1,2-diaminium dibromide (FcN2Br2)), which enables the use


of metallocenes in aqueous based RFBs (Figure 2)[40]. By coupling with methyl viologen (MV)
as the anolyte, this aqueous based organic flow battery achieved high theoretical energy density
of 45.5 Wh/L and excellent cycling performance from 40 to 100 mA/cm2. Notably, it also
demonstrated long cycling performance, 700 cycles at 60 mA/cm2 with 99.99% capacity

6
ACCEPTED MANUSCRIPT

retention, and delivered power density up to 125 mW/cm2. More recently, E.S. Beh et al.
reported a bis((3-trimethylammonio)propyl)ferrocene dichloride (BTMAP-Fc) as the catholyte
coupling with bis(3-trimethylammonio)propyl viologen tetrachloride (BTMAP-Vi) as the anolyte
in an aqueous based RFB, which has neutral pH, high capacity, high stability and low crossover
(Figure 3) [66]. They suggested that FcNCl is susceptible to decomposition via mechanisms that
involve the collision of two of the same molecules in a second-order process, while the BTMAP-

T
Fc can improve the stability especially at high concentration. They also discussed opportunities

IP
for future performance improvement, including chemical modification of a ferrocene center and
reducing the membrane resistance without significant increases in crossover. This approach may

CR
lead to cost-effective organic aqueous based RFBs for grid-scale electricity storage. Besides
ferrocene mentioned above, many other metallocenes also showed very promising potentials in

US
aqueous RFBs. For example, cobaltocene, nickelocne and molybdocenes all showed very
excellent electrochemical properties in both aqueous and non-aqueous solvents[67-69].
AN
M
ED
PT
CE
AC

Figure 2. (a) Preparation process for FcNCl and FcN2Br2. (b) cyclic voltammograms of FcNCl (red trace),
FcN2Br2 (purple trace), FcN (black trace), and MV (blue trace). (c) Extended 700 cycle testing data of the
0.5 M FcNCl/MV aqueous organic redox flow battery (AORFB) at 60 mA/cm2: capacity and coulombic
efficiency (CE) vs cycling numbers; (inset) representative charge and discharge profiles of selected cycles.
(d) polarization and power density curves of the FcNCl/MV AORFB at 0.5 (red traces) and 0.7 M (blue
traces) after full charge using 10 mA/cm2. Adapted from B. Hu et al., J. Am. Chem. Soc. 139, 1207-1214,
(2017)[40].

7
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Figure 3. (a) Chemical structures of BTMAP-Vi and BTMAP-Fc. (b) cyclic voltammograms of BTMAP-
Vi (blue trace) and BTMAP-Fc (red trace). (c) cycling of a BTMAP-Vi/BTMAP-Fc pH 7 cell at 50
mA/cm2. (d) representative voltage vs time traces of selected cycles. Adapted from E. Beh et al., ACS
Energy Lett. 2, 639-644, (2017)[66].
PT

Another type of metal-containing organic active materials is metal complexes. F. Burnea et al.
CE

exploited density functional theory (DFT) to investigate the tuning of reduction potential of the
first row transition metal (M = Cr, Mn, Fe, Co, Ni) complexes with the functionalized 1,4,7-
AC

Triazacyclononane-N,N',N”- triacetate (TCTA) ligands (Figure 4)[70]. They suggested 16


complexes as the active material in aqueous based RFBs, which can be classified into 5 groups
according to their potential ranges: Group I (-0.6V ~ -0.7 V), Group II (around 0.0V), Group III
(around 0.3 V), Group IV (0.6 ~ 0.8V) and Group V (1.1 ~1.2 V). Especially, it was suggested
that [MnLF], [MnLCN] and [NiLNH2] can be used as promising catholyte candidates owing to their
high reaction potentials reaching 1.25V in aqueous RFBs[70]. However, the estimation of

8
ACCEPTED MANUSCRIPT

solubility based on computational methods is yet a difficult task owing to lack of sufficient
experimental solubility data of similar compounds for verification. Thus, the practical
application of this type of metal-containing organic active materials is still under investigation.

T
IP
CR
US
AN
M

Figure 4. (a) Lowest Unoccupied Molecular Orbital (LUMO) of [Cr(1,4,7-Triazacyclononane-N,N',N”-


triacetate)] (Cr[TCTA]) (left) and [Fe(TCTA)] (center), and schematic figure for the structure of ligand
ED

functionalized complexes studied in this work (right) (b) calculated LUMO energy levels of spin ground
state for all neutral complexes and their absolute values with respect to that of each pristine complex
marked in parenthesis (eV). (c) The calibrated equilibrium adjusted reduction potential (E0) for all metal
PT

complexes ([MLNH2–MLNO2]) with respect to SHE and the controlled range by ligand modifications in V.
Adapted from F. Burnea et al., Electrochim. Acta. 246, 156-164, (2017)[70].
CE
AC

4.1.2. Metal-free organic active materials.

In aqueous based organic RFBs, the first demonstrated organic active material was tetrachloro-p-
benzoquinone (chloranil) as catholyte in a single flow acid Cd–chloranil battery in 2009 as
shown in Figure 5[71]. A high average coulombic efficiency of 99% and energy efficiency (EE)
of 82% over 100 cycles at the current density of 10 mA/cm2 was demonstrated. The same
research group further reported tiron (4, 5-dibenzoquione-1,3-benzenedisulfonate), which is a

9
ACCEPTED MANUSCRIPT

weak acid aromatic organic compound (a derivative of catechol), as an active material in


aqueous RFBs [72]. In this study, Y. Xu et al. investigated the electrochemical behavior of the
tiron and demonstrated the charge-discharge performance of a tiron/Pb test cell based on 3 M
H2SO4 electrolyte. Although this early work is only a semi-flow system, it provides a new class
of organic active materials for aqueous RFBs.

T
IP
CR
US
Figure 5. (a) Molecular structure and reaction of tetrachloro-p-benzoquinone. (b) cyclic
voltammograms the chloranil electrode. (c) a typical charge/discharge curve of the single flow acid Cd–
AN
chloranil battery at current density of 10 mA/cm-2. Adapted from Y. Xu et al., Electrochem. Commun. 11,
1422-1424, (2009)[71].
M

In 2014, the first quinone-based active material in aqueous RFBs has been demonstrated by
pairing with Br2/Br- redox couple and achieved a peak galvanic power density exceeding 0.6
ED

W/cm2 at 1.3 A/cm2 (Figure 6)[46]. The chemistry of 9,10-anthraquinone-2,7-disulphonic acid


(AQDS) underwent extremely rapid and reversible two-electron two proton reduction in sulfuric
PT

acid. The organic anthraquinone species can be synthesized from inexpensive commodity
chemicals, which represents a new and promising direction for realizing massive electrical
CE

energy storage at greatly reduced costs. In 2016, various anthraquinone derivatives have been
systematically investigated, and the anth-raquinone-2-sulfonic acid (AQS) which has a lower
AC

standard reduction potential than that of anthraquinone-2,7-disulfonic acid (AQDS) has been
demonstrated to have an increased peak galvanic power density over AQDS cells[47]. This study
demonstrated the effectiveness of using chemical synthesis to tailor organic molecules for
improving flow battery performance. S. Er et al. systematically studied the redox properties of a
library of existing and non-existing quinone derivatives and investigated their quantitative
structure-property relationships (QSPRs)[73]. They used a virtual screening approach coupled

10
ACCEPTED MANUSCRIPT

with materials genomic concepts to allow for the rational design of all-quinone flow batteries.
The identification of over 300 quinones with a predicted reduction potential above 0.7 V vs SHE
and the refined list of synthetically feasible quinones lay important foundation for achieving a
high-performance all-quinone flow battery.

T
IP
CR
US
AN
M
ED
PT

Figure 6. (a) Cell schematic. Discharge mode is shown; the arrows are reversed for electrolytic/charge
CE

mode. AQDSH2 refers to the reduced form of AQDS. (b) cyclic voltammogram of AQDS and DHAQDS
(1mM) in 1M H2SO4 on a glassy carbon electrode (scan rate=25 mV/s). (c) galvanic power density versus
current density for the same state of charge (SOC). (d) constant-current cycling at 0.5 A/cm2 at 40 °C
using a 3 M HBr + 0.5 M Br2 solution on the positive side and a 1 M AQDS + 1 M H2SO4 solution on the
AC

negative side. Adapted from B. Huskinson et al., Nature. 505, 195-198, (2014)[46].

Organic nitroxide radicals have been proved promising active materials in RFBs. TEMPO, with
four methyl groups adjacent to the N−O bond is one of the most commonly studied radicals[15,
74, 75]. TEMPO was first demonstrated in an aqueous based RFB coupling with viologen in
2015 with an excellent cycle life up to 10000 cycles (Figure 7)[76]. By using organic polymers

11
ACCEPTED MANUSCRIPT

as active material in combination with low-cost dialysis membranes, the system showed the
potential to realize a low-cost, safe and metal-free all-organic RFB. In an effort to increase the
energy density of TEMPO based RFBs, several TEMPO-based derivatives have been
synthesized and demonstrated in aqueous RFBs. T. Liu et al. modified TEMPO to a water-
soluble 4-hydroxy-2,2,6,6-tetramethylpiperidin-1-oxyl (4-HO-TEMPO) and demonstrated in an
all-organic aqueous flow battery with methyl viologen (MV) anolyte (Figure 8)[77]. The 4-HO-

T
TEMPO exhibited a high solubility up 2.1 M in water and delivered a stable capacity for 100

IP
cycles with nearly 100% coulombic efficiency. Additionally, the price of 4-HO-TEMPO based
aqueous RFBs was lower than $180/kWh versus $450/kWh for state of the art all vanadium flow

CR
batteries, approaching US Department of Energy (DOE) cost target on large-scale energy storage
of $150/kWh.

US
AN
M
ED
PT
CE
AC

Figure 7. (a) Electrolyte reactions of TEMPO radical and viologen. (b) time dependent NaCl
concentration of a chamber filled with deionized water that is separated from a NaCl feed solution (1
mol/L) by a dialysis membrane. Salt permeability was determined to be Ps = (9.3 ± 0.1) × 10-5 cm/s. (c) a

12
ACCEPTED MANUSCRIPT

representative cell voltage profile of test cell during constant-current cycling at 40 mA/cm2. (d) cycling
stability of the polymer-based RFB. Adapted from T. Janoschka et al., Nature. 527, 78-81, (2015)[76].

T
IP
CR
US
AN
M
ED
PT
CE
AC

Figure 8. (a) Cell reaction of the MV/4-HO-TEMPO aqueous based organic redox flow battery. (b) cyclic
voltammograms of MV (blue trace) and 4-HO-TEMPO (red trace). (c) capacity and coulombic efficiency
versus cycling numbers of the cell at 40 mA/cm2. Adapted from T. Liu et al., Adv. Energy Mater. 6,
1501449-1501457, (2016)[77].

Inspired from vitamin B2, K Lin et al. introduced a family of organic molecules alloxazine for
aqueous RFB applications (Figure 9)[58]. The alloxazine based active materials exhibits

13
ACCEPTED MANUSCRIPT

sufficiently high electrochemical and chemical stability with low reduction potential, which is
ideal to be exploited as an anolyte material in an alkaline RFB. Synthesis of the alloxazine based
active materials was carried out via facile coupling chemistry without employing high
temperature or pressure. These abundant materials can accelerate the development of organic
RFBs with low cost and high performance. The demonstrated current efficiency exceeded 99.7%
with a capacity retention of ~ 95% over 400 charge/discharge cycles cumulatively, or 99.98%

T
per cycle. We summarize the state-of-the-art organic active materials[58] in Table 1 together

IP
with the solubility, the electrolyte used and the demonstrated energy density.

CR
US
AN
M
ED
PT
CE

Figure 9. (a) Cyclic voltammogram of 2mM alloxazine 7/8-carboxylic acid (ACA) (red curve) and
ferrocyanide (gold curve) scanned at 100 mV/s on a glassy carbon electrode. (b) cell voltage and power
AC

density versus current density at 20 °C, at 10%, 50% and ~100% SOC. Electrolyte composition: 0.5 M
ACA and 0.4 M ferrocyanide + 40 mM ferricyanide were used in negative electrolyte and positive
electrolyte, respectively. (c) capacity retention, current efficiency and energy efficiency values over 400
cycles at 0.1 A cm-2. The normalized discharging capacity is evaluated on the basis of the capacity of the
first discharge cycle. (Inset: representative voltage versus time curves during 400 charge–discharge cycles
at 0.1 A cm-2, recorded between the first and fifth cycles.) Adapted from K. Lin et al., Nat. Energy. 1,
16102, (2016)[58].

14
ACCEPTED MANUSCRIPT

T
IP
Table 1. The key parameters of aqueous-based all-flow organic RFBs.
RFB types Cell voltage (V) Energy density (Wh/L) Performance Solubility in water Electrolyte Properties Year
> 100 cycle; 0.6
AQDS/Bromine[46, 78] ~ 0.858 ~ 16 AQDS ~ 1 M HBr and Br2; Acid-based 2014
W/cm2

CR
> 500 cycles; 60 No supporting electrolyte;
BTMAP-Vi/BTMAP-Fc[66] ~0.748 ~ 20 ~2M 2017
mW/cm2 Neutral-based
~ 25 cycles; ~250
DHDMBS/AQDS[79] < 0.6 < 10 ~2M 1 M H2SO4; Acid-based 2017
mW/cm2
> 400 cycles; ~
ACA/ferrocyanide[58] ~ 1.13 < 25 ACA ~ 1 M 1 M KOH; Alkaline-based 2016
350 mW/cm2
MV/4-HO-TEMPO[77]

2,6DHAQ/ferrocyanide[80]

TEMPTMA/MV[81]
~ 1.25

~ 1.20

~ 1.40
< 10

~ 6.8

~ 38
US
> 100 cycles; <
125 mW/cm2
> 100 cycles; >
450 mW/cm2
> 100 cycles; ~
250 mW/cm2
MV ~ 3 M;
4-HO-TEMPO ~ 2.1 M
2,6DHAQ > 0.6 M
TEMPTMA ~ 3.2 M; MV
~3M
1.5 M NaCl; Neutral-based

1 M KOH; Alkaline-based

1.5 M NaCl; Neutral-based


2016

2015

2016
AN
> 10000 cycles; ~ TEMPO < 0.1 M; Viologen
TEMPO/ Viologen[76] ~ 1.15 ~ 10 2 M NaCl; Neutral-based 2015
100 mW/cm2 < 0.1 M
~ 700 cycles; ~ FcNCl ~ 4 M; FcN2Br2 ~
FcNCl or FcN2Br2 /MV[40] ~ 1.0 ~ 10 2 M NaCl; Neutral-based 2017
125 mW/cm2 3.1 M; MV ~ 3 M
> 100 cycles; Quinoxaline ~ 4.5 M; BQ~ 1
Quinoxaline/BQ[28] ~ 1.4 < 10 M
Both Alkaline and acid 2015
N.G.
M
ED
PT
CE
AC

15
ACCEPTED MANUSCRIPT

4.2. Aqueous electrolyte


The relatively lower stability window of water compared to non-aqueous electrolyte is one of the
main drawbacks for aqueous based RFBs[82-84]. The design and operation of aqueous-based
RFBs shall always avoid or minimize water oxidation or reduction potentials. According to the
Pourbaix diagram of water, the stability region of water can be broadened by adjusting the pH in
the electrolyte. For acid electrolyte, the high solubility of proton can increase the ionic

T
conductivity of the electrolyte, which is critical to support organic molecule materials with fast

IP
reaction kinetics for realizing a higher power density. For example, the demonstrated AQS-
bromide flow cell in an acidic electrolyte (2 M H2SO4) achieved the power density of 0.7 W/cm2

CR
at 90% state of charge, which is the highest reported to date for a quinone-based flow battery at
room temperature (Figure 10)[85].

US
For alkaline electrolyte, K. Lin et al. reported highly soluble hydroxylated anthraquinones (AQ)
coupled with nontoxic ferricyanide ion (Figure 11) in an alkaline solution (1 M KOH) [80].
AN
They suggested that the OH group in the electrolyte are deprotonated to provide solubility while
acting as electron-donating groups to lower the reduction potential of the AQ molecule (i.e.
M

proton-dependent reactions). Different from the pH-dependent redox potential of AQ, the
ferro/ferricyanide redox potential is pH-independent, which allows the increase of open-circuit-
voltage (OCV) at high pH when coupling AQ with ferricyanide ion. The OCV of the reported
ED

alkaline quinone-based flow battery is 1.2 V at 50% SOC which is much higher than the acid
quinone-bromine flow battery at 0.8 V[80].
PT

Despite the advantages of high ionic conductivity, the acid or alkaline electrolytes suffer from
corrosion problems in cell stacks[86, 87]. Recently, more and more neutral pH based electrolytes
CE

are utilized in aqueous based organic RFBs. For example, the work of T. Liu et al. introduced the
neutral NaCl based supporting electrolyte for an aqueous based all organic RFB by using MV
AC

and 4-HO-TEMPO as active materials[77]. The use of low-cost and benign NaCl electrolyte
reduced the total capital cost of the RFB. The ionic conductivity can be improved by increasing
the Cl– charge carrier concentration. The work of E. Beh et al. using BTMAP-Vi and BTMAP-Fc
as active materials was also demonstrated in a neutral pH aqueous electrolyte. Since the active
materials consist of Cl– ions with high solubility (~2 M) in water, no additional supporting
electrolyte was added[66]. Their work demonstrated the highest capacity retention rate to date

16
ACCEPTED MANUSCRIPT

versus time and versus cycle number (Figure 3). It is clear that the pH selection of aqueous
electrolyte plays a crucial role in electrochemical kinetics, redox potential and reversibility of
aqueous organic based RFBs. To show the type of electrolytes used in combination of specific
redox molecules and their demonstrated performances, we included the electrolyte properties of
state-of-the-art aqueous based organic RFBs in Table 2.

T
IP
CR
US
AN
M
ED

Figure 10. (a) Pourbaix diagram of water. (b) cyclic voltammograms of AQDS (black), AQS (red),
DHAQDS (blue), and DHAQDMS (orange), showing the decrease in reduction potential through
chemical modification. (c) open circuit voltage as a function of state of charge for each quinone in a flow
PT

cell paired with the brominehydrobromic acid catholyte. (d) galvanic power density for AQS-bromide and
AQDS-bromide cells. Adapted from M. Gerhardt et al., Adv. Energy Mater. 7, 1601488-1601497,
(2017)[85].
CE
AC

17
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Figure 11. (a) Pourbaix diagram of 2,6-DHAQ. (b) molecular structures and CV of 1,5-DMAQ (olive),
plotted along with ferrocyanide (orange curve). (c) cell voltage and power density versus current density
ED

at 20° at 10, 50, and ~100% SOC. (d) representative voltage versus time curves during 100 charge-
discharge cycles at 0.1 A/cm2, recorded between the 10th and 19th cycles. Adapted from K. Lin et al.,
Science. 349, 1529-1532, (2015)[80].
PT

4.3. Membranes designs for aqueous organic RFBs


CE

Membrane is one of the most important and expensive components in RFBs[88-91]. An ideal
membrane should keep the anolyte and catholyte from mixing while still allowing transport of
AC

the charge carrying ions to maintain electrical neutrality and electrolyte balance[92-94]. The
most widely used aqueous based membrane is Nafion-based membrane[95]. The morphology of
Nafion is a hydrophobic and hydrophilic two-phase structure, in which the tetrafluoroethylene
(i.e., Teflon) forms the hydrophobic backbone while the sulfonate groups which are terminated
by pendant vinyl ether side chains exhibit hydrophilic property with proton conductivity [96].

18
ACCEPTED MANUSCRIPT

However, the cost of Nafion-based membranes contributes to the largest portion the cost of
RFBs[76, 97].

Generally, three main types of membranes have been utilized in the aqueous based organic RFBs
in addition to Nafion-based membranes. The first type is the microporous membrane. In 2015, T.
Janoschka et al. described an inexpensive dialysis membrane to separate the aqueous, polymer-
based anolyte and catholyte which can achieve current densities of up to 100 mA/cm2, and stable

T
long-term cycling capability (>10,000 cycles)[76]. The cellulose-based dialysis membrane in the

IP
work had a molecular-weight cut-off (MWCO; indicating the lowest retained molar mass) of

CR
6,000 g/mol. The redox-active polymers (TEMPO and Viol), which have a hydrodynamic radius
of around 2 nm, were effectively retained by the cellulose-based dialysis membrane with size-
exclusion, which has an estimated pore size of 1 nm. Minimum membrane selectivities (Smin) for

US
NaCl over the redox-active polymers (Permeability ratio: PNaCl / Ppolymer) were much higher than
the membrane selectivities of Nafion and some nanofiltration and microporous membranes[76].
AN
By combining simple dialysis membranes, which are only 5% to 10% of the cost of Nafion-
based membranes, the demonstrated RFB showed a promising affordable RFB concept (Figure
9). The microporous membrane is much cheaper compared with other ion exchange
M

membrane[98], which is a promising direction for developing lost-cost and high-performance


membranes for organic RFBs.
ED

Another type of membrane developed in aqueous based organic RFBs is anion exchange
membrane (AEM). B. Hu et al. investigated three ion exchange membranes (including Selemion
PT

AMV, ASV, DSV) for neutral aqueous organic RFBs[99]. The thinnest DSV membrane with the
lowest area resistance provided the best current dependent performance, energy efficiency,
CE

capacity utilization, and power density. The study emphasized that membrane resistance was an
important factor that determines the RFB performance. In another TEMPTMA/MV RFB system,
AC

a commercial fumasep FAA-3-PE-30 AEM has also been used to realize a stable cycle
performance[81]. These commercialized AEM showed very promising performance in aqueous
based organic RFBs. Additionally, Z. Chang et al. prepared a crosslinked methylated
polybenzimidazole (PBI) membrane which conducted the Cl- anion in the Zn/cation-grafted
TEMPO RFB, and achieved cycling stability over 140 cycles without any crossover, as shown in

19
ACCEPTED MANUSCRIPT

Figure 12[100]. However, the cell resistance was high (> 4 Ω), which could be attributed to the
thickness of the membrane and its ion-exchange ability.

The third type of membrane in aqueous based organic RFB is cation exchange membrane (CEM).
In 2017, J. Winsberg et al. demonstrated an aqueous Zn/2,2,6,6 Tetramethylpiperidine‑N‑oxyl
hybrid flow battery using CEM[101]. They reported that the quaternary ammonia species on the
pore surface of the AEMs could undergo a side reaction/precipitation of Zn2+, leading to

T
increasing cell resistance. A fumasep® F-930-RFD CEM, a membrane containing negatively

IP
charged sulfonic groups, was used in their system to repel the sulfonic group of the active

CR
material (TEMPO-4-sulfate). A. Orita et al. systematically investigated the membrane
performance in the Zn/ 4-hydroxy-2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPOL) hybrid flow
battery as shown in Figure 13[102]. Among the various combinations of membrane and

US
supporting electrolytes investigated, the use of a CEM with NaClO4 supporting electrolyte
showed the highest capacity retention, at 51% after 50 cycles. However, the active material,
AN
TEMPOL+, was found to react with CEM which could be the reason for capacity decay after
cycling. This suggests that investigating the chemical stability between membranes and active
materials is critical to ensure long-term stability of organic RFBs. Since the dissolved CO2 and
M

O2 in aqueous electrolyte can affect the chemical stability of the RFBs, the effect of air removal
from electrolytes was examined by N2-bubbling to the electrolyte[102]. The results showed that
ED

the capacity retention was improved more than 15% compared with non-bubbled system. Their
work provided a hint for further improving the membrane performance by adjusting pH and
PT

removing impurity from electrolyte[102].

Another strategy to decrease the cost of membrane is to develop the membrane-less RFBs. Some
CE

inorganic membrane-less flow battery systems have been proposed, including the soluble lead
acid[103], zinc–nickel[104], zinc–lead-dioxide [105] and zinc–cerium[106] systems. Many of
AC

these systems involve electrodeposition at one or both of their electrodes and take advantage of
the slow dissolution of the deposited metals in the presence of active species in the common
electrolytes. P. Leung et al. proposed a membrane-less hybrid flow battery by using soluble
organic quinone based catholyte with metallic zinc anolyte [107]. The dissolution rate of metallic
zinc was characterized in the presence of oxidized catholyte species as a mean to estimate
corrosion current. They showed that the highest corrosion current densities (1.9-9.4 mA/cm2) are

20
ACCEPTED MANUSCRIPT

lower than typical operating current density in RFBs (>30 mA/cm2) and concluded that a
membrane-less configuration is feasible for quinone based catholyte coupling with metallic zinc
in neutral aqueous electrolytes[107]. Recently, P. Navalpotro et al. reported a membrane-free
battery that relies on the immiscibility of organic molecule redox electrolytes (Figure 14)[108].
The proposed RFB has an OCV of 1.4 V with a high theoretical energy density of 22.5 Wh/L and
demonstrated 90% of its theoretical capacity with high stability (coulombic efficiency of 100%

T
and energy efficiency of 70%). This is the first demonstrated membrane-less organic molecule

IP
full flow battery system, providing a promising direction to develop a low-cost aqueous RFB
system.

CR
US
AN
M
ED
PT
CE
AC

Figure 12. (a) Synthesis route of a crosslinked and methylated meta-PBI membrane. (b) voltage profiles
of Zn - TEMPO RFB. (c) voltage efficiency versus current density. (d) cycling stability of the RFB.
Adapted from Z. Chang et al., ChemSusChem. 10, 3193-3197, (2017)[100].

21
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Figure 13. Cycling of Zn/TEMPOL RFBs using 0.1 M supporting electrolyte. 1st cycle voltage curves
from constant current charge/discharge at 10 mA cm-2, using (a) CEM and (b) an AEM; (c) discharge
capacities over 50 cycles. Adapted from A. Orita et al., J. Power Sources. 321, 126-134, (2016)[102].
PT
CE
AC

22
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN

Figure 14. (a) Schematic representation of the membrane-free battery concept based on immiscible redox
electrolytes. (b) Cyclic voltammetry experiments of 20 mM parabenzoquinone (pBQ) in 1-butyl-1-
M

methylpyrrolidinium bis(trifluoromethanesulfonyl)imide (PYR14TFSI) (green) and 20 mM H2Q in 0.1 M


HCl (blue) performed in 3-electrode electrochemical cells (scan rate=10 mV/s); (c) Discharge profiles of
the membrane-free battery at different current densities. (d) Cyclability study at 0.2 mA cm-2. coulombic
efficiency, energy efficiency, and capacity retention versus cycles. Adapted from P. Navalpotro et al.,
ED

Angew. Chem., 129, 12634 –12639 (2017)[108].


PT
CE
AC

23
ACCEPTED MANUSCRIPT

5. Non-aqueous based organic RFBs


One of the most critical bottlenecks of aqueous RFBs resides in their low energy density. The
energy density of RFBs is determined by both the cell voltage and the solubility of the active
materials dissolved in the catholyte and anolyte. Aqueous RFBs normally possess a lower energy
density than the non-aqueous RFBs because of their low operation voltage (below 2.0 V) limited
by water electrolysis. In fact, the low operation voltage bounded by water electrolysis wastes one

T
of the most important advantages of organic active material, i.e. wide reaction potential, as many

IP
of the high/low potential organic materials (e.g. [NiLNO2] (2.03 V vs. SHE), [CrLNH2] (-2.77 V
vs. SHE) exceed the stability window of the water electrolysis. To take full advantage of redox

CR
active organic materials with wide range of redox potentials, non-aqueous electrolytes are
employed in non-aqueous organic RFBs. However, the solubility of many organic materials in

US
non-aqueous solvents is typically less than 1.0 M, which corresponds to a volumetric energy
density less than 26.8 Ah/Lcatholyte. Much progress have been made to increase the concentration
AN
of organic RFBs. Here we review and categorize the organic redox active molecules employed in
non-aqueous systems into metal-containing and metal-free organic materials.
M

5.1. Active materials

5.1.1. Metal-containing organic active materials.


ED

The first attempt to utilize organic molecule as redox materials in non-aqueous RFBs was made
by Matsuda et al. in 1988, in which they used tris (2,2'-bipyridine) ruthenium(II) ([Ru(bpy)3]2+)
PT

complexes as active materials for both catholytes and anolytes. [Ru(bpy)3]2+ could be either
oxidized to [Ru(bpy)3]3+ at ~ 1.0 V vs Ag/Ag+, or reduced to [Ru(bpy)3]+ at -1.6 V vs Ag/Ag+
CE

which leads to an OCV of around 2.60 V[109]. Since then, a series of metal coordination
complexes with transitional metal center and various organic ligands has been evaluated in the
AC

non-aqueous RFBs.

Coordinated metal complexes

The coordinated metal complexes have many advantages over the bare metal ions. The redox
potential of the metal coordination complex is mainly determined by the metal center.
Accompanied with the newly introduced organic ligands, the redox potential of the metal

24
ACCEPTED MANUSCRIPT

coordination complex can be tuned by changing the interactions between the organic ligands and
the transitional metal center. Generally, the redox potential of the metal center will be lowered
with the addition of electron-donating ligands; the redox potential of the metal center will be
elevated with the addition of electron-withdrawing ligands. Other properties such as solubility,
stability, kinetics can also be tailored to meet design needs by modifying the organic ligands.
The early developed coordinated metal non-aqueous RFB systems are mainly based on anion-

T
exchange systems[109-115], which means the redox couples for both the anolyte and catholyte

IP
are based on the same metal cation center with different oxidation state. Their oxidation states
change with the operation of the cell coupled with the net anion exchange in the electrolyte.

CR
Hence, AEM was normally adopted to separate the redox cation couples with different oxidation
state[109-116]. Although these systems enjoy high theoretical cell voltage (> 2V), they suffer

US
from low energy efficiency, poor cycle life and limited solubility. The performance of
representative non-aqueous RFBs based on coordinated metal complex are summarized in Table
2.
AN
M

Metallocene
ED

Metallocenes, represents another group of organometallic molecules with the metal ions
sandwiched with two aromatic cyclopentadienyl rings[117]. The metal ion redox center can
undergo one electron transfer and the redox potential for this to happen is still mainly determined
PT

by the metal center. One of the most reported metallocenes is ferrocene. Ferrocene has long been
used as quasi-reference redox molecules for its excellent stability, superior reversibility and
CE

relatively stable redox potential in different solvents[118]. And with the flexibility brought by
the two cyclopentadienyl rings, researchers could easily be able to modify this sandwiched
AC

molecule for different purposes[13, 119, 120]. Ding et al. reported the high rate ferrocene
cathode in 2015, and by coupling with the pretreated lithium metal, a membrane-free system was
designed and a rate as high as 60 C-rate was achieved. However, this system is still limited by
the poor solubility of ferrocene in the carbonate based electrolyte (0.6 M)[42]. By the
substitution of one acetyl group on the cyclopentadienyl ring of ferrocene, Kim et al.
demonstrated the acetyl ferrocene (AcFc) catholyte with concentration of 0.8 M in propylene

25
ACCEPTED MANUSCRIPT

carbonate (PC) based electrolyte[120]. Wei et al. further improved the solubility of an ionic
ferrocene compound, ferrocenylmethyl dimethyl ethyl ammonium
bis(trifluoromethanesulfonyl)imide (Fc1N112-TFSI) to 1.7M in the pristine carbonate mixture of
ethylene carbonate (EC)/PC/ethyl methyl carbonate (EMC) (4:1:5 by weight), and 0.85M in
EC/PC/EMC with 1.2 M LiTFSI. Nuclear magnetic resonance (NMR) and DFT calculation
revealed that the increased solubility of Fc1N112-TFSI in carbonate is mainly due to the

T
enhanced solvation interactions between the solvent and the newly introduced polar motif[13].

IP
Later, Cong et al. showed that the solubility of ferrocene derivative could be further improved in
carbonate based electrolytes. By adding one methyl group to each of the cyclopentadienyl ring,

CR
the symmetry of the parent ferrocene molecule is broken, and the melting point of the resulted 1,
1-dimethylferrocene (DMFc) is significantly reduced to 37 – 40 °C, marginally higher than room

US
temperature. The solubility of DMFc is also greatly improved, and become nearly immiscible
with pristine diethyl carbonate (DEC) [119].
AN
Despite all the encouraging improvement in the ferrocene-based catholytes, the performance of
the lithium metal anode still hinders the development of the ferrocene system. Another inherent
drawback of the lithium metal anode is its poor scalability compared with the ferrocene based
M

catholytes. A more promising alternative option is realized by replacing lithium metal with
anolytes containing low redox potential molecules[121, 122]. Taking advantage of the superior
ED

electrochemical properties of the metallocence, Hwang et al. first demonstrated an all


metallocene RFB with ferrocene catholyte and cobaltocene (Cc) anolyte, and by introducing
PT

electron-withdrawing bromide to ferrocene and ten electron-donating methyl groups to


cobaltocene, an expanded cell voltage of 2.0 V was achieved in acetonitrile[121]. Later, Ding et
CE

al. also investigated the Ferrocene/cobaltocene system by optimizing the organic solvents. With
the application of an solid LISICON-type electrolyte to prevent the mixing of 1,3-dioxolane
(DOL) in the cobaltocene anolyte and N,N-dimethylformamide (DMF) in the ferrocene catholyte
AC

they successfully increased the cell voltage from 1.3 V when using the same solvent to 1.8
V[122]. Another problem of the all-metallocene system is its relative low cell voltage compare
with the lithium metal anode system. To fully exploit the superior electrochemical properties of
the metallocene catholyte, redox couples with high solubility, good reversibility and most
importantly low potential must be developed. For instance, aromatic hydrocarbons can be
reversibly reduced to aromatic radical anions at the potential as low as 0.5 V vs Li/Li+[123]. In

26
ACCEPTED MANUSCRIPT

addition, Yu et al reported a liquid sodium biphenyl anode having a low reaction potential as low
as 0.09 V vs Na/Na+, which was coupled with polysulfide catholyte and demonstrated more than
3500 cycles[124]. Therefore, we suggest that a full flow RFB with cell voltage higher than 3.0 V
could be designed by using this low redox potential biphenyl as anolyte and ferrocene derivatives
as catholyte. We summarized the representative metallocene based non-aqueous RFBs in Table
3.

T
IP
CR
US
AN
M
ED
PT
CE
AC

27
ACCEPTED MANUSCRIPT

T
IP
Table 2. Summary of coordinated metal complexes used in non-aqueous RFBs. Modified and adapted from [125].

Redox couple Redox Demonstrated Electrolyte Membrane Current Energy Performance Year Ref.
concentration density

CR
potential (V) (M) collector (Wh/L)

2+ 3+ +
[Ru(bpy)3] /[Ru(bpy)3] 1.0 V vs Ag/Ag 0.02 0.1M Et4NBF4 AEM Glassy 0.134 4 cycles 1988 [109]
in ACN carbon (GC)
+
[Ru(bpy)3] /[Ru(bpy)3] 2+
-1.6 V vs Ag/Ag +
0.02 plate 40% CE

[Fe(bpy)3]2+/[Fe(bpy)3]3+

[Ni(bpy)3]/[Ni(bpy)3]2+
0.65 V vs Ag/Ag+

-1.66 V vs Ag/Ag+
0.4

0.2
0.5M TEABF4
in PC US
AEM
GC plate

Carbon felt

Carbon felt
20.84 91.4% CE, 81.8%
EE, 5 cycles
2012 [113]
AN
2+ 3+ +
[V(acac)3] /[V(acac)3] 0.45V vs Ag/Ag 0.01 0.5M TEABF4 AEM Graphite <0.2 50% CE at 50% 2009 [112]
in ACN SOC, 2 cycles
+ 2+ +
[V(acac)3] /[V(acac)3] -1.75 V vs Ag/Ag 0.01 Graphite

[Cr(acac)3]/[Cr(acac)3]+ 1.20V vs Ag/Ag+ 0.05 0.5M TEABF4 AEM Graphite <1 53-58% CE, 21- 2010 [111]
in ACN 22% EE at 50%
M

[Cr(acac)3]-/[Cr(acac)3] -2.20 V vs Ag/Ag+ 0.05 Graphite SOC, 5 cycles

+ +
[Mn(acac)3]/[Mn(acac)3] 0.7V vs Ag/Ag 0.05 0.5M TEABF4 AEM Graphite <0.4 21% EE, 74-97% 2011 [115]
in ACN CE, 2 cycles
- +
[Mn(acac)3] /[Mn(acac)3] -0.4 V vs Ag/Ag 0.05 Graphite
ED
PT
CE
AC

28
ACCEPTED MANUSCRIPT

T
IP
Table 3. Summary of metallocene molecules used in non-aqueous RFBs.

Redox couple Redox Demonstrated Electrolyte Membrane Current Energy Performance Year Ref.
Concentration density

CR
potential (V) (M) collector (Wh/L)

+ +
Fc/Fc 3.25 V vs Li/Li 0.6 1 M LiPF6 in Celgard Carbon paper 40 500 cycles 2015 [42]
EC:DEC
Li metal anode 99.96%capacity
retention per cycle

AcFc/AcFc+

Li metal anode
3.66 V vs Li/Li+ 0.81 1 M LiPF6 in PC
US
Celgard/Nafion
117
Carbon paper 79.2 >99% CE, 400 cycles 2016 [120]
AN
Fc1N112-TFSI 3.49V vs Li/Li+ 0.85 1M LiTFSI in Celgard Graphite felt 50 75% EE, 20 cycles for 2014 [13]
EC/PC/EMC flow test
Li metal anode

DMFc/DMFc+ 3.1V vs Li/Li+ 3 3M LiClO4 in Celgard and Carbon paper 204 >95% CE, 40 cycles 2017 [119]
EC:DEC LAGP
Li metal anode
M

Fc/Fc+ 0.041V vs Ag/Ag+ 0.01 1M TEAPF4 in AEM Graphite <1 78.6% CE, 3 cycles 2014 [121]
ACN
+ +
Cc/Cc -1.29 V vs Ag/Ag 0.01 Graphite
ED

BrFc/BrFc+ 0.219V vs Ag/Ag+ 0.01 1M TEAPF4 in AEM Graphite <1 82.1% CE, 3 cycles 2014 [121]
-1.83 V vs Ag/Ag+ ACN
Cc*/Cc*+ 0.01 Graphite

+ +
Fc/Fc 0.5V vs Ag/Ag 0.1 0.5 M LiPF6 in LATP Ti foil coated ~5 99% capacity retention 2017 [122]
PT

DMF with per cycle, >95%


Cc/Cc +
-1.9Vvs SCE 0.1 acetylene CE, >85% EE, 30
0.5M LiTFSI in black cycles
DOL
CE
AC

29
ACCEPTED MANUSCRIPT

5.1.2. Metal-free organic active materials.


Different from metal-containing organic active materials where the redox reaction is mainly
realized by the valence change of the metal center, the redox reaction of metal-free organic
material is based on the change of the charge state of the organic moiety with conjugated
structure or atoms with lone pair electrons such as S, O and N. A comprehensive review paper
on solid organic materials by Song et al.[16] grouped most of the reported organic molecules into

T
seven categories based on the structure of the organic redox center: (Table 4): conjugated

IP
hydrocarbon, conjugated amine, conjugated thioether, organosulfide, thioether, nitroxyl radical
and conjugated carbonyl[124]. Alternatively, the organic electrode materials can also be

CR
categorized into three groups based on the redox reactions (Figure 15). N-type organic materials
could shift their charge state between neutral state (N) and negatively charged state (N-) (e.g.

US
Organosulfide and conjugated carbonyl); while p-type organic materials could shift their charge
state between neutral state and positively charged state (e.g. conjugated amine and conjugated
AN
thioether). The last one is bipolar organic materials of which their neutral state can be either
reduced to negatively charged state and oxidized to positively charged state (e. g. conjugated
hydrocarbon and nitroxyl radical). Next, we discuss each type of structure with representative
M

examples and their applications in non-aqueous organic RFBs.


ED
PT
CE
AC

Figure 15. Reaction mechanism of three type of elctroactive organic materials. (a) n-type; (b) p-type; (c)
bipolar. Adapted from J. Song et al., Energy Environ. Sci., 6, 2280–2301 (2013)[8].

30
ACCEPTED MANUSCRIPT

Conjugated hydrocarbon/conjugated amine/conjugated thioether

Conjugated hydrocarbon, conjugated amine and conjugated thioether were first studied as
conducting polymers. MacDiarmid et al. first studied polyacetylene (PAc) as cathode material
for batteries and since then many conductive polymers were investigated for the application of
solid electrode materials[126]. Basically, all these polymers could be categorized into five

T
typical groups: polyparaphenylene (PPP)[127, 128], polyacetylene (PAc)[126, 127], polyaniline
(PAn)[129, 130], polypyrrole (PPy)[131, 132] and polythiophene (PTh)[133, 134]. Since the

IP
solubility of all these conductive polymers are very low in either aqueous or non-aqueous based

CR
electrolytes, all the studies of these polymers were aimed to employ them as solid electrode
material[126, 127, 129, 130, 132-134], and chemical modifications were usually taken to
reduced their solubility.

US
For the application of redox-active molecules in RFBs, high solubility is a very critical criterion,
AN
since the solubility of the active material directly determines the energy density of the RFB. Two
directions could realize the use of these conducting polymers in the non-aqueous RFBs. The first
one is to increase their solubility by introducing functional groups to the parent molecule, aiming
M

to enhance the interactions between the organic solvent and the modified polymer molecule[76].
The second one is to employ the recent developed concept of semisolid-suspension-based
ED

electrolyte by Duduta et al.[135], which breaks the solubility limitation of the redox active
molecules in RFBs. For instance, a conductive slurry could be fabricated by mixing the solvent
PT

with conducting polymer/carbon composite and used as flowing catholyte or anolyte. The first
polymer-based non-aqueous symmetric RFB was reported by Oh et al. in 2014[136], in which
they prepared an suspension-based electrolyte of PTh particles. PTh servers as the bipolar redox-
CE

active material and a cell voltage of 2.5 V was achieved. The performance of the PTh suspension
was tested in a flow cell setup with AEM as separator. The flow battery demonstrated a stable
AC

charge/discharge behavior over 20 cycles, and an energy efficiency of 61% was achieved. Other
promising redox active polymers based on Organosulfides, Thioethers, Nitroyl radicals and
Conjugated carbonyls could also be used in the non-aqueous RFBs with this strategy. Next we
will focus on the discussion of small molecules of the aforementioned organics.

31
ACCEPTED MANUSCRIPT

Organosulfides

The redox centers of organosulfides including organodisulfides[137-139] and


organotrisulfide[140, 141] reside on the S atom, and two/four-electron transfer happens with the
reversible broken and rebuilt of the S-S bond in organodisulfide and organotrisulfide

T
respectively. Visco et al. first reported the application of tetraethylthiuram disulfide (TETD) as

IP
cathode material for high temperature sodium batteries[139]. Later they screened a series of

CR
other dimeric organosulfides for solid lithium batteries[137, 138]. Recently Wu et al. reported
the use of dimethyl trisulfide (DMTS)[141] and diphenyl trisulfide (DPTS) [140]as active
material for catholyte. Three sulfur atoms allow four electron storage per molecule affording

US
DMTS and DPTS with a theoretical capacity of 849 and 428 mAh/g respectively[140, 141].
However, the most prominent problem of this system is that one of the discharge product of the
AN
trisulfides catholyte is Li2S which barely soluble in most non-aqueous solvents.

Organodisulfide polymers were also investigated for the use of solid electrode material in the
M

1990s[142]. Generally, all the organodisulfides can be divided into two categories: organosulfide
polymer with S-S bond in the main chain[143-147], organosulfide polymer with S-S bond in the
ED

side chain[148, 149]. After discharge, the long chain of the former polymer is cut into small
lithiated molecules while for the latter polymer the backbone remains intact. Although the
solubility of the polymeric organodisulfides in most organic solvent are very limited, they all
PT

have potential application in the semi-solid based RFBs as mentioned in the earlier section.
CE

Nitroxyl radical

Nitroxyl radicals belong to the bipolar organics, but most of the researches on nitroxyl radical
AC

focus on their higher potential redox process. Stable nitroxyl radicals usually have alkyl or aryl
substituents to the N atom. As one of the most stable radicals, TEMPO has four methyl groups
adjacent to the N-O bond offering steric protection of the radical center. A variety of TEMPO
based derivatives have been investigated for their application in the non-aqueous based RFBs
because of their superior stability and reversibility[15, 150]. Because of high dipole moment of
~3.0D[43], TEMPO has a low melting point (36-38 ºC) and excellent solubility in polar organic

32
ACCEPTED MANUSCRIPT

solvents. All these facts together with the its redox potential at ~3.5V vs Li/Li+ make TEMPO a
promising materials for non-aqueous RFBs. Wei et al. studied the TEMPO catholyte in 2014[15].
The solubility of TEMPO in EC/PC/EMC mixture is as high as 5.2 M, and remains to be 2.0 M
when 2.3M LiPF6 is introduced. By coupling with lithium-graphite anode, the 2.0 M TEMPO
catholyte delivered an impressive discharge energy density of 126 Wh/L, which is almost five-
time that of aqueous all-vanadium flow battery[15]. Moreover, the TEMPO catholyte has also

T
been evaluated by coupling with other liquid anolyte, e.g. TEMPO/N-methylphthalimide[74] and

IP
Camphoquinone/4-oxo-TEMPO[75]. Takechi et al. studied the TEMPO derivative 4-methoxy-2,
2, 6, 6-tetra-methylpiperidine-1-oxyl (MeO-TEMPO) as cathode material, and demonstrated that

CR
a concentration higher than 2.5M is impossible to obtain with traditional routes to prepare the
catholyte[150]. They found that Meo-TEMPO forms a deep eutectic solvent free catholyte (DES)

US
with the addition of LiTFSI. The resulted DES remained in the liquid state over a broadened
temperature range, and delivered a stable capacity as high as 200Wh/L[150].
AN
Conjugated carbonyl compounds

The redox mechanism of the carbonyl compound can be generalized as an enolization reaction
M

and a reverse reaction of the carbonyl group[16]. Stabilizing the chemically reactive radical is
crucial in order to apply the carbonyl compound in organic RFBs. Wei et al. studied the 9-
ED

fluorenone anolyte coupled with a DBBB derivative catholyte and obtained a non-aqueous all
organic RFB[51]. The spin and charge in the reduced 9-fluorenone (FL•-) were resonated onto
PT

the adjacent aromatic ring. They also found that supporting electrolyte and solvents also played a
crucial role in the stabilization of FL•-[51].
CE

As another well-known reprehensive carbonyl molecule, quinones show greatly improved


stability for 2e- processes, which is promoted by the two conjugated carbonyl groups. The 2e-
AC

reduction processes of quinones are stepwise in non-aqueous electrolyte with the formation of
radical anion followed by dianion. Although quinones have been used for solid electrode
materials for a long time, their application in the non-aqueous RFBs have seldom been studied
until recent years. Wang et al. first studied an organic/inorganic hybrid RFBs based on an
anthraquinone derivative with methoxytriethyleneglycol substituents (15D3GAQ) dissolved in
PC as catholyte[151]. The stepwise reactions of quinones in non-aqueous were clearly showed

33
ACCEPTED MANUSCRIPT

by the two sets of redox peaks in cyclic voltammograms and two-plateaus in the galvanostatic
charge and discharge curves. The ethylene glycol moieties enhanced the solubility of 15D3GAQ
in non-aqueous polar solvent to a concentration of at least 0.25 M. The static cell demonstrated a
theoretical specific discharge energy density of 25 Wh/L with an energy efficiency of 82% over
nine cycles. Noticeable capacity fading caused by side reactions between anthraquinone and the
carbonate solvents was observed, thus further optimization of the electrolyte is necessary to

T
improve the performance of this system[151].

IP
CR
US
AN
M
ED
PT
CE
AC

34
ACCEPTED MANUSCRIPT

T
IP
Table 4. Summary of representative redox molecules used in non-aqueous RFBs. Modified and adapted from [16].

Structure Redox mechanism Examples demonstrated in non- Ref.


aqueous RFBs

CR
Conjugated [126-128]
hydrocarbons

Conjugated amine
US [129-132]
AN
Conjugated thioether [133, 134]
M

Organosulfide [140, 141]


DPTS DMTS

Nitroxyl radical [15, 43, 77,


ED

100]

TEMPO Meo-TEMPO

Conjugated carbonyl [51, 144]


PT
CE

FL
AC

35
ACCEPTED MANUSCRIPT

5.2. Non-aqueous electrolytes

Here we discuss key design considerations of non-aqueous electrolyte for organic RFBs
including solubility, electrochemical window, and viscosity.

Solubility

T
While the solubility of active materials depends on both the active materials and the solvent,

IP
some classes of non-aqueous solvents intrinsically possesses higher solvation ability than other
non-aqueous solvents. Typically, the solubility of a solute particle is determined by the dynamic

CR
interaction between the solute and the surrounding solvent molecules[152, 153]. The solvent
molecules in the solvation sphere constantly interchange with those in the bulk and the ratio of

US
the free solvent molecules to the ones in the solvation sphere could be used to characterize the
solute-solvent interactions. When the ratio decreases, the solute-solvent interaction becomes
more intense and vice versa. The solubility of organic molecules could be estimated by the
AN
Hansen solubility parameters (HSP) in which the solute-solvent interaction is simplified into
three major types: nonpolar interactions (often referred as dispersion interaction, associated with
M

van der Waals forces), polar interaction (associated with dipole moments) and hydrogen bond. A
good HSP match between the solute and solvent always indicates a decent solubility[154, 155].
ED

For instance, as a promising candidate, ferrocene attracts researchers with its excellent stability,
superior reversibility and fast kinetics. However, the solubility of ferrocene is too low for the
application in RFBs. The low solubility of ferrocene in most commonly used organic solvents
PT

could be explained by the mismatch between the highly nonpolar ferrocene molecule and the
polar nature of the solvent. By introducing an ionic charged tetraalkylammonium pendant arm
CE

with a TFSI− counter anion to the cyclopentadienyl ring of ferrocene, Wei et al. improved the
solubility of Fc1N112-TFSI to 1.7M in the pristine carbonate mixture of EC/PC/EMC (4:1:5 by
AC

weight), and 0.85M in the same carbonate mixture with 1.2 M LiTFSI. The origin of the
increased solubility has been investigated NMR and DFT calculations, and the underlying reason
is attributed to the enhanced solvation interactions between the solvent and the newly introduced
polar motif[13].

36
ACCEPTED MANUSCRIPT

The solubility of salt in organic solvent can also be roughly predicted by the solvent’s dielectric
constant which describes its relative chemical polarity. Following the rules of “like dissolves
like”[152, 153], solvents with high dielectric constant can effectively improve the salt
dissociation and solvate the resulted cation and anion. The aforementioned solute-solvent
interactions still come into play in the salt dissolving process.

T
IP
Electrochemical stability window

CR
The most prominent advantage of the non-aqueous RFBs over aqueous RFBs is the greatly
expanded electrochemical window (ECW), which is practically defined as the potential range
between the cathodic and anodic decomposition limit of the solvent/supporting electrolyte
US
system. The ECW of most organic solvents used in non-aqueous RFBs is greater than 3.0 V,
which is much higher than that of aqueous electrolytes (~1.23 V). Compared with the anodic
AN
limit, the cathodic limit of the solvent is more critical when one try to enlarge the cell voltage by
lowering the redox potential of the anode material. Several parameters substantially influence the
M

determined ECWs of organic solvents including supporting electrolytes, electrode materials, and
temperature. The ECWs of most commonly used organic solvents are depicted in Figure 16,
with test conditions briefly elucidated. Among these solvents, Sulfolane (SL) has the lowest
ED

cathodic potential limit (-4.0 V vs. SCE) and a sufficient anodic potential limit (2.3 V vs. SCE);
however, its application in battery technology is limited by its high viscosity, low ionic
PT

conductivity and relatively high freezing point (28.4 ºC)[156]. As a compromise, solvents with
low viscosity and freezing points are often used to mix with SL to lower the freezing point and
CE

reduce the viscosity[102, 157]. Ionic liquids (ILs) have been emerging as a promising alternative
for solvents due to their intrinsic ionic conductivity, wide potential window and low saturated
AC

vapor pressure, however they suffer from high viscosity. The strategy to use binary or ternary
solvent mixture could be applied to ILs to further extend their application in RFBs.

37
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Figure 16 The electrochemical windows of most commonly used organic solvents in non-aqueous RFBs.
Adapted from O. Luca et al., Org. Chem. Front., 2, 823-848 (2015)[156].
ED
PT

Viscosity

Viscosity determines the ionic conductivities of solvents as governed by strokes law as follows:
CE
AC

Where is the limiting molar conductivity, is the charge number, is the Faraday constant,
is the Avogadro constant, is the Strokes constant, is the viscosity of the solvent and is
the ionic radius. In order to reduce the internal cell resistance and hence improve the energy
efficiency, organic solvents with low viscosity are preferred. In addition, solvent viscosity also
positively related to the pump loss. Fluids with higher viscosity lead to larger pump loss at the
same flowing rate. Although the viscosity generally increases with the increasing concentration

38
ACCEPTED MANUSCRIPT

of solute (including active material and salt)[150], tuning the solvent property such as
hydrophobicity using highly concentrated salt could serve as strategy to reduce the semi-solid
type of electrolyte[158]. A tradeoff between the energy density (partially determined by active
material concentration) and the energy efficiency (related to cell ohmic resistance of the cell,
pump loss etc.) must be studied and considered in optimizing the electrolyte composition. Other
parameters such as dielectric constant, toxicity, saturated vapor pressure etc. are also equally

T
important in the choice of solvents for non-aqueous RFBs in practical applications.

IP
CR
5.3. Membranes designs for non-aqueous organic RFBs

US
Membranes play a much more critical role in conventional all liquid RFBs than they do in
batteries with solid electrodes. Unlike the solid electrode, the redox couples in liquid phase will
freely “crossover” driven by either concentration gradient or electric field. The ideal membrane
AN
will keep the redox couples from mixing while allow the supporting ions passing through to
balance the charge. Generally, there are two classes of membranes applied in RFBs: organic
M

polymer and inorganic ceramics. For aqueous RFBs, polymer membranes such as Nafion are
often used for their good stability, high ionic conductivity and excellent flexibility[5, 88-90, 96,
159, 160]. However, their application in non-aqueous based electrolyte is limited because of the
ED

high permeability of redox couples caused by membrane swelling due to serious solvent uptake
(especially in non-aqueous electrolytes), which consequently enlarges the pore size and reduces
PT

the selectivity of the membrane. This problem can be addressed by doping the polymeric
membranes with organic-solvent-resistant materials such as polyvinylidene difluoride (PVDF).
CE

Jia et al.[161] mixed Nafion and PVDF with different ratio and they found that selectivity of the
newly fabricated membrane was greatly improved. However, the resistivity of this membrane
AC

was also increased. Following this method, other organic-solvent-resistant material with better
ionic conductivity such as poly (ethylene oxide) (PEO) may be adopted for this method to
increase the ionic conductivity of the hybrid polymeric membrane for the application in non-
aqueous based RFBs.

To date, the most reported membranes used in non-aqueous and aqueous/non-aqueous hybrid
RFBs are inorganic ceramics[119, 150]. The first study on the ionic conductivity of the inorganic

39
ACCEPTED MANUSCRIPT

ceramics can date back to 1830s, when Faraday found the remarkable ionic conductivity of Ag2S
and PbF2 at high temperature[162]. However, the first application of ion conducting ceramic in
batteries was reported in 1960s when β-alumina (Na2O·11Al2O3) was found to have excellent
sodium ion conductivity and was subsequently used in high temperature sodium sulfur
batteries[27, 97]. Since then, the sodium conductive β-alumina has been incorporated various
other battery systems[2, 163] and the β-alumina based high-temperature sodium-sulfur battery

T
has already been commercialized in Japan[14]. Inorganic lithium ion conductive ceramics was

IP
developed and used in lithium-ion battery to avoid the potential fire or explosion dangers caused
by overcharging or short-circuit in 1990s. Since then inorganic lithium-ion conductive ceramics

CR
such as perovskite-type, NASICON-type, and garnet-type have been vigorous studied.

Perovskites
US
With a general formula of AMO3, the Li+ transportation is realized by the absorption of Li+ into
AN
the A-deficient-site first and then migrates to either distorted tetrahedral sites or the square planar
face of the MO3 that does not face a A3+ ion[164]. As a typical perovskite-type ceramic lithium
ion conductor, the ionic conductivity of lithium lanthanum titanate (LLTO) Li3xLa0.67−xTiO3
M

(0<x<0.16) was reported to be as high as 10-3 S/cm at ambient temperature[164], however this
type of material contains TiIV which will be reduced when it is in contact with metallic lithium.
ED

Sodium (Na) super ionic conductor (NASICON) type


PT

NASICON-type materials generally have a formula of AM2(PO4)3 where A represent Li, Na or K


and M could be Ge, Zr or Ti[18]. As one of the widely studied NASICON type material,
CE

LiTi2(PO4)3 was found to have very low ionic conductivity, which was later improved by doping
LiTi2(PO4)3 with M (M=Al, Ga, Cr, Fe, Sc, In, Lu Y or La) to form Li1+xMxTi2-x(PO4)3[95, 165].
Among them, the Al substituent (LATP) has demonstrated the highest ionic conductivity[94,
AC

166]. A significant drawback of this material is its instability against metallic lithium caused by
the reduction of Ti4+[26, 32]. With Zr4+ confined in an isolated octahedron, LiZr2(PO4)3 (LZP)
was considered to be more stable and investigated as a replacement of LATP[39, 154]. The same
strategy of doping element used in LATP to increase its ionic conductivity equally apply to LZP
and its ionic conductivity could be further optimized by controlling the doping content.

40
ACCEPTED MANUSCRIPT

Garnets

Garnet-type materials have a cubic A3B2C3O12 structure with A and B cations have eight-fold
and six-fold coordination respectively. Li-rich garnets have been investigated extensively since
1969 because of their high Li+ conductivity (10-4 S/cm) at room temperature as well as superior
stability against lithium metal[155]. However, its degradation caused by H+ and Li+ exchanging

T
was observed when they are in contact with atmospheric moisture. Improving the density of
garnet ceramics by sintering condition control was often adopted[19].

IP
In general, inorganic ion conducting ceramics are superior to the polymeric membranes for the

CR
application in non-aqueous based RFBs. However, the one critical drawback of these inorganic
ion conductors is their fragile nature. Development of flexible composite membranes[19, 154,

US
155], which exploit both merits of polymeric membrane and ceramic conductor is a promising
strategy to achieve high-performance and low-crossover membranes for non-aqueous organic
AN
RFBs.
M

6. Perspective

Organic RFBs are promising technologies owing to its unique advantages in low cost and diverse
ED

material selection. However, the development is still in its early stage. To further improve
organic RFBs for both stationary and automobile applications, we here discuss remaining critical
PT

challenges and suggest approaches to address these challenges (Figure 17).


CE
AC

41
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Figure 17. Summary of the perspective of organic redox flow batteries. (a) Strategies to further increase
the energy density of organic RFBs; (b) Strategies to expand the operation voltage of organic RFBs; (c)
design considerations for membrane development beyond Nafion; (d) critical needs in modeling work to
ED

guide the cell and stack design optimized for practical applications.

(1). Increase the volumetric capacity and energy density. The volumetric capacity of the
state-of-art vanadium RFBs is ≈ 50 Ah/L resulting in energy density ≈ 25 Wh/L[5, 24]. However,
PT

the demonstrated volumetric capacity for most of the organic RFBs is ≈ 10 Ah/L, resulting in
low energy density 10 Wh/L. The low volumetric capacity is associated with the low
CE

concentration/solubility of the organic active materials. The search for organic active species
with higher solubility is important. For example, 4-HO-TEMPO and AQS exhibit an appreciable
AC

solubility of ≈2.1 and 2.0 M in water, respectively[77] [47, 85], which is comparable to
vanadium ions. To date, the demonstrated highest energy density of aqueous based organic all
flow batteries is 38 Wh/L in a TEMPTMA/MV system in 2016 owing to high solubility of
TEMPTMA (2.3 M) and MV (2.4 M)[81]. Modifying the functional group on the organic
molecules and tuning the composition of the supporting electrolyte (solvent/salt) are two
effective approaches to further improve the solubility of the organic molecules[13, 119, 120]. For

42
ACCEPTED MANUSCRIPT

instance, adding of two methyl groups on the ferrocene molecule increases the solubility from
less than 1.0 M to more than 3.0M[119]. Kim et al. demonstrated the acetyl ferrocene catholyte
with concentration of 0.8 M in PC based electrolyte by the substitution of one acetyl group on
the cyclopentadienyl ring of ferrocene [120].

In addition to increase the solubility, the semi-solid slurry approach[135] represents another
effective way to improve the volumetric capacity of RFBs. This approach was first reported for

T
intercalation type of chemistry by M. Duduta et al.[135]. By mixing insoluble and non-

IP
conducting solid active materials with conducting carbon in organic electrolyte, the demonstrated

CR
semi-solid suspension breaks the limitation of active material solubility and achieved much
higher energy density than all soluble active material based RFBs (Figure 17). Thereafter, many
other insoluble Li-ion based active species (sulfur[158, 167, 168], polysulfide[10, 169],

US
silicon[170, 171], LFP[172, 173], LTO[174] etc.) have been demonstrated as the semi-solid
suspensions in non-aqueous based RFBs. In 2013, the first aqueous based semi-solid flow battery
AN
has been proposed by Z. Li et al. utilizing LTP and LFP as anolyte and catholyte[175]. Similarly,
one could apply the semi-solid approach for organic active species to increase the volumetric
capacity. For example, many organic materials designed for the application of solid electrode
M

and shuttle additives used for Li-ion battery over-charge protection have a high reaction potential
and good cycle stability, however they have low solubility (<0.5 M) in both aqueous and non-
ED

aqueous electrolytes. Fabricating organic based semi-solid suspension will address their
solubility issue and achieve low-cost and high-energy organic RFBs. We propose and summarize
PT

several potential organic materials for semi-solid suspension-based RFBs, as shown in Table 5.
Another novel way to increase the energy density of the organic RFBs is realized by redox
CE

targeting methods proposed by Wang’s group[161, 173, 176]. Multiple redox mediators serve as
charge carriers exchanging electrons between the electrodes (electrochemical reactions) and high
energy density solid phase active materials (chemical reactions) stored in the reservoir.
AC

43
ACCEPTED MANUSCRIPT

T
IP
Figure 16. (a) Scheme of semi-solid flow cell (SSFC) system using flowing lithium-ion cathode and
anode suspensions. (b) Fluid semi-solid suspension containing LiCoO2 powder as the active material and

CR
Ketjen black as the dispersed conductive phase, dispersed in alkyl carbonate electrolyte. (c) Galvanostatic
charge/discharge curves for semi-solid suspensions having 26 vol% LiCoO2 (LCO) dispersed in 1.3 M
LiPF6 in an alkyl carbonate blend. Adapted from M. Duduta et al., Adv. Energy Mater. 1, 511-516,
(2011)[135].

US
AN
Table 5. The key parameters of potential organic materials for semi-solid suspensions.
Materials Reaction Potential (V) Specific Capacity (mAh/g) Cycle Performance Function
PTO[177] 0.51 vs. SHE 395 > 1500 cycles Aqueous-based anode
M

PPTO[177] -0.07~0.04 vs. SHE 144 ~ 229 > 1000 cycles Aqueous-based anode
PAQS[177] -0.6 vs. SHE 200 > 1000 cycles Aqueous-based anode
MPT[178] 3.47 vs. Li/Li+ ~ 120 > 200 cycles Redox shuttle
PFPTFBB[179] 4.3 vs. Li/Li+ N.G. > 100 cycles Redox shuttle
ED

TEDBPDP[180] 4.8 vs. Li/Li+ N.G. ~ 10 cycles Redox shuttle


THHQ[181] 3.5 vs. Li/Li+ 628 ~ 40 cycles Li-ion cathode
LC[182] 3.5 vs. Li/Li+ ~ 169 ~ 50 cycles Li-ion cathode
HATN-CMP[183] 2.8 vs. Li/Li+ ~ 150 ~ 50 cycles Li-ion cathode
PT

(2). Enlarge the voltage window of organic RFBs. There is large room to improve the stable
CE

potential window of both aqueous and non-aqueous based electrolytes considering the diverse
potential window of organic active materials. For aqueous organic RFBs, the cell voltage is
limited by the water stability window. Searching for suitable redox pairs that make the most use
AC

of the water stable potential window is crucial. The average working potential for VRBs is ≈1.25
V[25], which is higher than the state-of-the-art aqueous organic RFBs (< 1.0 V)[31]. According
to the Pourbaix diagram of water, the oxygen reduction potential increases with decreasing pH
and the hydrogen evolution potential decreases with increasing pH. Using hybrid electrolyte
system could significantly improve the effective cell voltage of aqueous RFBs.[160, 184]
Another way to enlarge the voltage window of aqueous organic RFBs is to apply highly-
44
ACCEPTED MANUSCRIPT

concentrated electrolytes[185] including water-in-salt (WiS)[153] and hydrate melt aqueous


electrolytes[156]. We believe that this new class of highly concentrated electrolyte will be
effective to enlarge the potential window of aqueous organic RFBs.

For non-aqueous organic RFBs, metallic lithium was usually adopted as anode to provide high
cell voltage, which however reduced the scalability of the cell and limited the cycle life. Nearly
all the active materials used in organic RFBs are concentrated on the potential range between 2.0

T
to 3.8 V vs Li/Li+, which strongly limits the advantage of high stability at wide potential window

IP
of non-aqueous solvents. The search for highly soluble organic redox couple with low potential
(< 1.5 V vs Li/Li+) represents promising direction for the development of future non-aqueous

CR
based organic RFBs[8, 122]. For example, aromatic hydrocarbons were reported to have stable
redox potential as low as 0.5V vs Li/Li+[123]. Further characterization and optimization of

US
organic molecules with low redox potentials are critical to replace the currently used problematic
metallic lithium. With the help of well-developed theoretical computing methods, such as DFT,
AN
researchers are calculating the redox potentials of thousands of molecules as an effective
screening method before synthesis[46, 53, 108, 186, 187]. The pace of designing new low
potential organic molecules will be greatly accelerated in the near future.
M

(3). Develop low-cost and high-performance membranes. Nafion-based membranes are the
ED

state-of-the-art membrane used for aqueous based organic RFBs. However, the high-cost of
Nafion-based membrane account for almost 40% of the cost of the total stack [80, 159]. The
development of low-cost and high-performance membranes is one of the most important
PT

challenges for all RFBs. For instance, T. Liu et al. reported a low-cost AEM in their total organic
aqueous RFB employing methyl viologen anolyte and 4-HO-TEMPO catholyte[77]. They
CE

estimated the cell cost is lower than $180/kWh versus $450/kWh for state of the art all vanadium
RFBs owing to low-cost membranes. To further increase the ionic conductivity, improve the
AC

selectivity and reduce the membrane cost for organic RFBs, we believe that approaches and
lessons learned from conventional RFBs are valuable directions including microporous
membrane with size screening selection, asymmetric pore size distribution, and cation or anion
charge-repulsion etc. [97, 152, 153, 188, 189]. In non-aqueous based organic RFBs, much
serious electrolyte organic solvent uptake usually makes the polymeric membrane swollen which
consequently enlarges the pore size and reduces the selectivity of the membrane.[161] To date,

45
ACCEPTED MANUSCRIPT

the most reported membranes used in non-aqueous RFBs are inorganic ceramics [119, 150].
However, their fragile nature caused by high elastic moduli is still an obvious drawback for their
application in RFBs. Development of flexible composite membranes, which exploit both merits
of polymeric membrane and ceramic conductor, seems more promising[190, 191]. In addition, it
is also an important issue to investigate the reactivity/compatibility between membranes and
organic active materials in both aqueous and non-aqueous RFBs[8, 161, 186]. Finally, the

T
membrane structure and requirement highly depend on the type and charge-storage mechanism

IP
of the active materials and should be design as a system.

CR
(4). Stack and system design and modeling. There has been substantial development of cell,
stack and system design for VRFBs [162, 192-194]. However, there are very few existing studies
focusing on the cell, stack and system design and testing for organic RFBs. These studies are

US
critical for scale-up applications. Modeling and simulation studies on organic RFBs are mostly
reported on the single cell level[195, 196]. For example, D. Chu et al. developed an enhanced
AN
non-isothermal transient model for a metal-free quinone–bromide flow battery in a single cell
configuration and concluded that the flow rate has little effect on cell performance when low
current density is applied[197]. At a stack level, additional considerations such as electrolyte
M

distribution inside the cell stack [198-200] and shunt current issue should be studied. In
conventional RFBs, many flow frame design[194] and cell interaction investigation[8, 19] have
ED

been carried out to address these issues. In addition, the use of ceramic membrane in organic
RFBs will add new complexity to the cell stack design due to the brittleness of ceramics. In short,
PT

while many existing modeling and simulation methods reported for the cell stack design for
VRFBs[201, 202] could serve as important foundation for organic RFBs, new design constraints
CE

and considerations associated with the unique characteristics of organic RFBs also present new
challenges and opportunities for developing large-scale organic RFB systems.
AC

(5). Long-Term Cycling Life. There is still lack of large-scale long-term demonstration for both
aqueous and non-aqueous based organic RFBs. One of the most important targets is to realize
stable organic RFB systems that offer long-term cycling life, in both cycle number and cycle
time. To date, existing aqueous based organic RFBs (e.g. TEMPO-viologen system [76]; FcNCl-
MV system [40]) have achieved more than 1000 cycles. The high cycling number is achieved
owing to stable redox couples and effective membranes that alleviate the crossover. In contrast,

46
ACCEPTED MANUSCRIPT

the cycling stability of non-aqueous based organic RFBs is still low due to limited selection of
effective membranes and chemical compatibility issues with non-aqueous electrolytes. For
example, very few organic polymer membranes can be used in non-aqueous based organic RFBs
due to the chemical stability with the non-aqueous solvents. Ceramic membranes are typically
applied at the expense of ionic conductivity. In addition to cycle numbers, it is also critical to
investigate the time-dependent cell degradation behaviors [66] or shelf life issues due to

T
chemical degradation. In short, the long-term cycling stability of organic RFBs is a result of

IP
multiple factors (e.g. chemical and electrochemical reversibility of active materials, selectivity of
the membranes, stability of electrode and electrolyte etc.) and shall be addressed at each level

CR
from active materials to holistically at the system level.

Acknowledgment US
This work is supported by a grant from the Research Grants Council (RGC) of the Hong Kong
AN
Administrative Region, China, under Theme-based Research Scheme through Project No. T23-
60I/17-R.
M

References
ED

[1] International Energy Outlook 2016, With Projections To 2040; U.S. Energy Information
Administration (EIA), DOE/EIA-0484 (2016).
[2] E. Zanzola, C.R. Dennison, A. Battistel, P. Peljo, H. Vrubel, V. Amstutz, H.H. Girault,
Electrochim. Acta, 235 (2017) 664-671.
PT

[3] Renewables 2017 Global Status Report; REN21.


[4] Z.G. Yang, J.L. Zhang, M.C.W. Kintner-Meyer, X.C. Lu, D.W. Choi, J.P. Lemmon, J. Liu,
Chem. Rev., 111 (2011) 3577-3613.
CE

[5] M. Skyllas-Kazacos, M.H. Chakrabarti, S.A. Hajimolana, F.S. Mjalli, M. Saleem, J.


Electrochem. Soc., 158 (2011) R55-R79.
[6] K. Wang, K. Jiang, B. Chung, T. Ouchi, P.J. Burke, D.A. Boysen, D.J. Bradwell, H. Kim, U.
Muecke, D.R. Sadoway, Nature, 514 (2014) 348-350.
AC

[7] S. Roe, C. Menictas, M. Skyllas-Kazacos, J. Electrochem. Soc., 163 (2016) A5023-A5028.


[8] Y. Zhao, Y. Ding, Y. Li, L. Peng, H.R. Byon, J.B. Goodenough, G. Yu, Chem. Soc. Rev., 44
(2015) 7968-7996.
[9] B. Li, Z. Nie, M. Vijayakumar, G. Li, J. Liu, V. Sprenkle, W. Wang, Nat. Commun., 6 (2015)
6303-6311.
[10] W. Wang, Q. Luo, B. Li, X. Wei, L. Li, Z. Yang, Adv. Funct. Mater., 23 (2013) 970-986.
[11] C. Menictas, M. Skyllas-Kazacos, J. Appl. Electrochem., 41 (2011) 1223-1232.
[12] A. Manthiram, Y. Fu, Y.-S. Su, J. Phys. Chem. Lett., 4 (2013) 1295-1297.

47
ACCEPTED MANUSCRIPT

[13] X. Wei, L. Cosimbescu, W. Xu, J.Z. Hu, M. Vijayakumar, J. Feng, M.Y. Hu, X. Deng, J.
Xiao, J. Liu, Adv. Energy Mater., 5 (2015) 1400678.
[14] A.N. Colli, P. Peljo, H.H. Girault, Chem. Commun., 52 (2016) 14039-14042.
[15] X. Wei, W. Xu, M. Vijayakumar, L. Cosimbescu, T. Liu, V. Sprenkle, W. Wang, Adv.
Mater . , 26 (2014) 7649-7653.
[16] Z. Song, H. Zhou, Energy Environ. Sci., 6 (2013) 2280-2301.
[17] K. Gong, Q. Fang, S. Gu, S.F.Y. Li, Y. Yan, Energy Environ. Sci., 8 (2015) 3515-3530.
[18] J. Rugolo, M.J. Aziz, Energy Environ. Sci., 5 (2012) 7151-7160.
[19] G.-M. Weng, Z. Li, G. Cong, Y. Zhou, Y.-C. Lu, Energy Environ. Sci., 10 (2017) 735-741.

T
[20] E. Sum, M. Rychcik, M. Skyllas-Kazacos, J. Power Sources, 16 (1985) 85-95.
[21] E. Sum, M. Skyllas-Kazacos, J. Power Sources, 15 (1985) 179-190.

IP
[22] M. Skyllas-Kazacos, M. Rychick, R. Robins, in, Google Patents, 1988.
[23] M. Ulaganathan, V. Aravindan, Q. Yan, S. Madhavi, M. Skyllas‐Kazacos, T.M. Lim, Adv.

CR
Mater. Interfaces, 3 (2016) 1500309.
[24] F. Rahman, M. Skyllas-Kazacos, J. Power Sources, 189 (2009) 1212-1219.
[25] L. Li, S. Kim, W. Wang, M. Vijayakumar, Z. Nie, B. Chen, J. Zhang, G. Xia, J. Hu, G.
Graff, J. Liu, Z. Yang, Adv. Energy Mater., 1 (2011) 394-400.

Lett., 13 (2013) 1330-1335. US


[26] B. Li, M. Gu, Z. Nie, Y. Shao, Q. Luo, X. Wei, X. Li, J. Xiao, C. Wang, V. Sprenkle, Nano

[27] C. Xie, Y. Duan, W. Xu, H. Zhang, X. Li, Angew. Chem. Int. Ed., 56 (2017) 14953-14957.
AN
[28] F.R. Brushett, A.N. Jansen, J.T. Vaughey, L. Su, J.D. Milshtein, in, Google Patents, 2015.
[29] G.L. Soloveichik, Chem. Rev., 115 (2015) 11533-11558.
[30] X. Wei, W. Pan, W. Duan, A. Hollas, Z. Yang, B. Li, Z. Nie, J. Liu, D. Reed, W. Wang,
ACS Energy Lett., 2 (2017) 2187-2204.
M

[31] P. Leung, A. Shah, L. Sanz, C. Flox, J. Morante, Q. Xu, M. Mohamed, C.P. de León, F.
Walsh, J. Power Sources, 360 (2017) 243-283.
[32] J. Winsberg, T. Hagemann, T. Janoschka, M.D. Hager, U.S. Schubert, Angew. Chem. Int.
ED

Ed., 56 (2017) 686-711.


[33] S. Evers, T. Yim, L.F. Nazar, J. Phys. Chem. C, 116 (2012) 19653-19658.
[34] H. Pan, X. Wei, W.A. Henderson, Y. Shao, J. Chen, P. Bhattacharya, J. Xiao, J. Liu, Adv.
Energy Mater., 5 (2015) 1500113-1500120.
PT

[35] Y. Yang, G.Y. Zheng, Y. Cui, Chem. Soc. Rev., 42 (2013) 3018-3032.
[36] K.B. Hatzell, M. Boota, Y. Gogotsi, Chem. Soc. Rev., 44 (2015) 8664-8687.
[37] P. Leung, X. Li, C.P. De León, L. Berlouis, C.J. Low, F.C. Walsh, RSC Adv., 2 (2012)
CE

10125-10156.
[38] T.B. Schon, B.T. McAllister, P.-F. Li, D.S. Seferos, Chem. Soc. Rev., 45 (2016) 6345-6404.
[39] K. Wedege, E. Dražević, D. Konya, A. Bentien, Sci. Rep., 6 (2016) 39101.
AC

[40] B. Hu, C. DeBruler, Z. Rhodes, T.L. Liu, J. Am. Chem. Soc., 139 (2017) 1207-1214.
[41] Y. Zhao, Y. Ding, J. Song, G. Li, G. Dong, J.B. Goodenough, G. Yu, Angew. Chem. Int.
Ed., 53 (2014) 11036-11040.
[42] Y. Ding, Y. Zhao, G. Yu, Nano Lett., 15 (2015) 4108-4113.
[43] J. Lalevee, X. Allonas, P. Jacques, Comput. Theor. Chem., 767 (2006) 143-147.
[44] W. Duan, R.S. Vemuri, J.D. Milshtein, S. Laramie, R.D. Dmello, J. Huang, L. Zhang, D. Hu,
M. Vijayakumar, W. Wang, J. Mater. Chem. A, 4 (2016) 5448-5456.
[45] W. Duan, R.S. Vemuri, D. Hu, Z. Yang, X. Wei, J. Vis. Exp., (2017) e55171-e55171.

48
ACCEPTED MANUSCRIPT

[46] B. Huskinson, M.P. Marshak, C. Suh, S. Er, M.R. Gerhardt, C.J. Galvin, X. Chen, A.
Aspuru-Guzik, R.G. Gordon, M.J. Aziz, Nature, 505 (2014) 195-198.
[47] B. Yang, L. Hoober-Burkhardt, F. Wang, G.S. Prakash, S. Narayanan, J. Electrochem. Soc.,
161 (2014) A1371-A1380.
[48] F.R. Brushett, J.T. Vaughey, A.N. Jansen, Adv. Energy Mater., 2 (2012) 1390-1396.
[49] J.D. Milshtein, L. Su, C. Liou, A.F. Badel, F.R. Brushett, Electrochim. Acta, 180 (2015)
695-704.
[50] W. Duan, J. Huang, J.A. Kowalski, I.A. Shkrob, M. Vijayakumar, E. Walter, B. Pan, Z.
Yang, J.D. Milshtein, B. Li, ACS Energy Lett., 2 (2017) 1156-1161.

T
[51] X. Wei, W. Xu, J. Huang, L. Zhang, E. Walter, C. Lawrence, M. Vijayakumar, W.A.
Henderson, T. Liu, L. Cosimbescu, Angew. Chem. Int. Ed., 54 (2015) 8684-8687.

IP
[52] J. Huang, L. Cheng, R.S. Assary, P. Wang, Z. Xue, A.K. Burrell, L.A. Curtiss, L. Zhang,
Adv. Energy Mater., 5 (2015) 1401782.

CR
[53] G. Gryn'Ova, J.M. Barakat, J.P. Blinco, S.E. Bottle, M.L. Coote, Chem.-Eur. J, 18 (2012)
7582-7593.
[54] L. Zhang, Z. Zhang, K. Amine, Redox Shuttle Additives for Lithium-Ion Battery, in:
Lithium Ion Batteries-New Developments, InTech, 2012.

US
[55] C.S. Sevov, D.P. Hickey, M.E. Cook, S.G. Robinson, S. Barnett, S.D. Minteer, M.S.
Sigman, M.S. Sanford, J. Am. Chem. Soc., 139 (2017) 2924-2927.
[56] H. Huang, R. Howland, E. Agar, M. Nourani, J.A. Golen, P.J. Cappillino, J. Mater. Chem.
AN
A, 5 (2017) 11586-11591.
[57] Geological Survey Vanadium Mineral Commodities Summary (2016),
https://minerals.usgs.gov/minerals/pubs/mcs/2016/mcs2016.pdfUSUS
[58] K. Lin, R. Gómez-Bombarelli, E.S. Beh, L. Tong, Q. Chen, A. Valle, A. Aspuru-Guzik, M.J.
M

Aziz, R.G. Gordon, Nat. Energy, 1 (2016) 16102.


[59] C.S. Sevov, S.L. Fisher, L.T. Thompson, M.S. Sanford, J. Am. Chem. Soc., 138 (2016)
15378-15384.
ED

[60] Y. Liang, Z. Tao, J. Chen, Adv. Energy Mater., 2 (2012) 742-769.


[61] Z. Chen, Y. Qin, K. Amine, Electrochim. Acta, 54 (2009) 5605-5613.
[62] G. Kear, A.A. Shah, F.C. Walsh, Int. J. Energy Res., 36 (2012) 1105-1120.
[63] J. Noack, N. Roznyatovskaya, T. Herr, P. Fischer, Angew. Chem. Int. Ed., 54 (2015) 9776-
PT

9809.
[64] Y. Wang, P. He, H. Zhou, Adv. Energy Mater., 2 (2012) 770-779.
[65] P. Van sek, aynes, W. M., Ed. CRC Press/Taylor and Francis Boca Raton, FL, (2014).
CE

[66] E.S. Beh, D. De Porcellinis, R.L. Gracia, K.T. Xia, R.G. Gordon, M.J. Aziz, ACS Energy
Lett., 2 (2017) 639-644.
[67] S. Gubin, S. Smirnova, L. Denisovich, A. Lubovich, J. Organomet. Chem., 30 (1971) 243-
255.
AC

[68] W.E. Geiger Jr, W.L. Bowden, N. El Murr, Inorg. Chem., 18 (1979) 2358-2361.
[69] K.L. Breno, T.J. Ahmed, M.D. Pluth, C. Balzarek, D.R. Tyler, Coordin. Chem. Rev., 250
(2006) 1141-1151.
[70] F.K.B. Burnea, H. Shi, K.C. Ko, J.Y. Lee, Electrochim. Acta., 246 (2017) 156-164.
[71] Y. Xu, Y. Wen, J. Cheng, G. Cao, Y. Yang, Electrochem. Commun., 11 (2009) 1422-1424.
[72] Y. Xu, Y.-H. Wen, J. Cheng, G.-P. Cao, Y.-S. Yang, Electrochim. Acta, 55 (2010) 715-720.
[73] S. Er, C. Suh, M.P. Marshak, A. Aspuru-Guzik, Chem. Sci., 6 (2015) 885-893.

49
ACCEPTED MANUSCRIPT

[74] Z. Li, S. Li, S. Liu, K. Huang, D. Fang, F. Wang, S. Peng, Electrochem. Solid State Lett., 14
(2011) A171-A173.
[75] S.-K. Park, J. Shim, J. Yang, K.-H. Shin, C.-S. Jin, B.S. Lee, Y.-S. Lee, J.-D. Jeon,
Electrochem. Commun., 59 (2015) 68-71.
[76] T. Janoschka, N. Martin, U. Martin, C. Friebe, S. Morgenstern, H. Hiller, M.D. Hager, U.S.
Schubert, Nature, 527 (2015) 78-81.
[77] T. Liu, X. Wei, Z. Nie, V. Sprenkle, W. Wang, Adv. Energy Mater., 6 (2016) 1501449.
[78] B. Huskinson, M. Marshak, M. Gerhardt, M.J. Aziz, in, The Electrochemical Society, 2014,
pp. 27-30.

T
[79] L. Hoober-Burkhardt, S. Krishnamoorthy, B. Yang, A. Murali, A. Nirmalchandar, G.S.
Prakash, S. Narayanan, J. Electrochem. Soc., 164 (2017) A600-A607.

IP
[80] K. Lin, Q. Chen, M.R. Gerhardt, L. Tong, S.B. Kim, L. Eisenach, A.W. Valle, D. Hardee,
R.G. Gordon, M.J. Aziz, Science, 349 (2015) 1529-1532.

CR
[81] T. Janoschka, N. Martin, M.D. Hager, U.S. Schubert, Angew. Chem. Int. Ed., 55 (2016)
14427-14430.
[82] M. Vijayakumar, L. Li, G. Graff, J. Liu, H. Zhang, Z. Yang, J.Z. Hu, J. Power Sources, 196
(2011) 3669-3672.

Sci., 4 (2011) 4068-4073. US


[83] W. Wang, S. Kim, B. Chen, Z. Nie, J. Zhang, G.-G. Xia, L. Li, Z. Yang, Energy Environ.

[84] J. Zhang, L. Li, Z. Nie, B. Chen, M. Vijayakumar, S. Kim, W. Wang, B. Schwenzer, J. Liu,
AN
Z. Yang, J. Appl. Electrochem., 41 (2011) 1215-1221.
[85] M.R. Gerhardt, L. Tong, R. Gómez-Bombarelli, Q. Chen, M.P. Marshak, C.J. Galvin, A.
Aspuru-Guzik, R.G. Gordon, M.J. Aziz, Adv. Energy Mater., 7 (2017) 1601488.
[86] H. Liu, Q. Xu, C. Yan, Y. Qiao, Electrochim. Acta, 56 (2011) 8783-8790.
M

[87] H. Liu, Q. Xu, C. Yan, Electrochem. Commun., 28 (2013) 58-62.


[88] Q. Luo, H. Zhang, J. Chen, P. Qian, Y. Zhai, J. Membrane. Sci. , 311 (2008) 98-103.
[89] Q. Luo, H. Zhang, J. Chen, D. You, C. Sun, Y. Zhang, J. Membrane. Sci., 325 (2008) 553-
ED

558.
[90] X. Li, P. Vandezande, I.F. Vankelecom, J. Membrane Sci., 320 (2008) 143-150.
[91] W. Lu, Z. Yuan, Y. Zhao, L. Qiao, H. Zhang, X. Li, Energy Storage Materials, 10 (2018)
40-47.
PT

[92] A. Parasuraman, T.M. Lim, C. Menictas, M. Skyllas-Kazacos, Electrochim. Acta, 101 (2013)
27-40.
[93] Y. Zhao, Z. Yuan, W. Lu, X. Li, H. Zhang, J. Power Sources, 342 (2017) 327-334.
CE

[94] S. Gu, K. Gong, E.Z. Yan, Y. Yan, Energy Environ. Sci., 7 (2014) 2986-2998.
[95] D. Reed, E. Thomsen, W. Wang, Z. Nie, B. Li, X. Wei, B. Koeppel, V. Sprenkle, J. Power
Sources, 285 (2015) 425-430.
[96] K. Schmidt-Rohr, Q. Chen, Nat. Mater., 7 (2008) 75-83.
AC

[97] D. Chen, D. Li, X. Li, J. Power Sources, 353 (2017) 11-18.


[98] H. Prifti, A. Parasuraman, S. Winardi, T.M. Lim, M. Skyllas-Kazacos, Membranes, 2 (2012)
275-306.
[99] B. Hu, C. Seefeldt, C. DeBruler, T.L. Liu, J. Mater. Chem. A, 5 (2017) 22137-22145.
[100] Z. Chang, D. Henkensmeier, R. Chen, ChemSusChem, 10 (2017) 3193-3197.
[101] J. Winsberg, C. Stolze, A. Schwenke, S. Muench, M.D. Hager, U.S. Schubert, ACS Energy
Lett., 2 (2017) 411-416.
[102] A. Orita, M. Verde, M. Sakai, Y. Meng, J. Power Sources, 321 (2016) 126-134.

50
ACCEPTED MANUSCRIPT

[103] D. Pletcher, R. Wills, Phys. Chem. Chem. Phys., 6 (2004) 1779-1785.


[104] J. Cheng, L. Zhang, Y.-S. Yang, Y.-H. Wen, G.-P. Cao, X.-D. Wang, Electrochem.
Commun., 9 (2007) 2639-2642.
[105] J. Pan, Y. Sun, J. Cheng, Y. Wen, Y. Yang, P. Wan, Electrochem. Commun., 10 (2008)
1226-1229.
[106] P. Leung, C. Ponce-de-Leon, C. Low, A. Shah, F. Walsh, J. Power Sources, 196 (2011)
5174-5185.
[107] P. Leung, T. Martin, A. Shah, M. Anderson, J. Palma, Chem. Commun., 52 (2016) 14270-
14273.

T
[108] P. Navalpotro, J. Palma, M. Anderson, R. Marcilla, Angew. Chem. Int. Ed., 56 (2017)
12460-12465.

IP
[109] Y. Matsuda, K. Tanaka, M. Okada, Y. Takasu, M. Morita, T. Matsumura-Inoue, J. Appl.
Electrochem., 18 (1988) 909-914.

CR
[110] M.H. Chakrabarti, E.P.L. Roberts, C. Bae, M. Saleem, Energy Convers. Manag., 52 (2011)
2501-2508.
[111] Q. Liu, A.A. Shinkle, Y. Li, C.W. Monroe, L.T. Thompson, A.E. Sleightholme,
Electrochem. Commun., 12 (2010) 1634-1637.

11 (2009) 2312-2315. US
[112] Q. Liu, A.E. Sleightholme, A.A. Shinkle, Y. Li, L.T. Thompson, Electrochem. Commun.,

[113] J. Mun, M.-J. Lee, J.-W. Park, D.-J. Oh, D.-Y. Lee, S.-G. Doo, Electrochem. Solid State
AN
Lett., 15 (2012) A80-A82.
[114] A.A. Shinkle, A.E. Sleightholme, L.T. Thompson, C.W. Monroe, J. Appl. Electrochem.,
41 (2011) 1191-1199.
[115] A.E. Sleightholme, A.A. Shinkle, Q. Liu, Y. Li, C.W. Monroe, L.T. Thompson, J. Power
M

Sources, 196 (2011) 5742-5745.


[116] T. Yamamura, Y. Shiokawa, H. Yamana, H. Moriyama, Electrochim. Acta, 48 (2002) 43-
50.
ED

[117] D. Astruc, New J. Chem., 33 (2009) 1191-1206.


[118] R.R. Gagne, C.A. Koval, G.C. Lisensky, Inorg. Chem., 19 (1980) 2854-2855.
[119] G. Cong, Y. Zhou, Z. Li, Y.-C. Lu, ACS Energy Lett., 2 (2017) 869-875.
[120] H.-s. Kim, T. Yoon, Y. Kim, S. Hwang, J.H. Ryu, S.M. Oh, Electrochem. Commun., 69
PT

(2016) 72-75.
[121] B. Hwang, M.S. Park, K. Kim, ChemSusChem, 8 (2015) 310-314.
[122] Y. Ding, Y. Zhao, Y. Li, J.B. Goodenough, G. Yu, Energy Environ. Sci., 10 (2017) 491-
CE

497.
[123] N. Holy, Chem. Rev., 74 (1974) 243-277.
[124] J. Yu, Y.-S. Hu, F. Pan, Z. Zhang, Q. Wang, H. Li, X. Huang, L. Chen, Nat. Commun., 8
(2017) 14629.
AC

[125] F. Pan, Q. Wang, Molecules, 20 (2015) 20499-20517.


[126] P.J. Nigrey, D. MacInnes, D.P. Nairns, A.G. MacDiarmid, A.J. Heeger, J. Electrochem.
Soc., 128 (1981) 1651-1654.
[127] L. Shacklette, J. Toth, N. Murthy, R. Baughman, J. Electrochem. Soc., 132 (1985) 1529-
1535.
[128] L. Zhu, A. Lei, Y. Cao, X. Ai, H. Yang, Chem. Commun., 49 (2013) 567-569.
[129] A. MacDiarmid, L. Yang, W. Huang, B. Humphrey, Synth. Met., 18 (1987) 393-398.
[130] N. Gospodinova, L. Terlemezyan, Prog. Polym. Sci., 23 (1998) 1443-1484.

51
ACCEPTED MANUSCRIPT

[131] M. Zhou, J. Qian, X. Ai, H. Yang, Adv. Mater . 23 (2011) 4913-4917.


[132] N. Mermilliod, J. Tanguy, F. Petiot, J. Electrochem. Soc., 133 (1986) 1073-1079.
[133] L. Liu, F. Tian, X. Wang, Z. Yang, M. Zhou, X. Wang, React. Funct. Polym., 72 (2012)
45-49.
[134] K. Kaneto, K. Yoshino, Y. Inuishi, Jpn. J. Appl. Phys., 22 (1983) L567-L568.
[135] M. Duduta, B. Ho, V.C. Wood, P. Limthongkul, V.E. Brunini, W.C. Carter, Y.-M. Chiang,
Adv. Energy Mater., 1 (2011) 511-516.
[136] S. Oh, C.-W. Lee, D. Chun, J.-D. Jeon, J. Shim, K. Shin, J. Yang, J. Mater. Chem. A, 2
(2014) 19994-19998.

T
[137] M. Liu, S.J. Visco, L.C. De Jonghe, J. Electrochem. Soc., 138 (1991) 1891-1895.
[138] M. Liu, S.J. Visco, L.C. De Jonghe, J. Electrochem. Soc., 137 (1990) 750-759.

IP
[139] S.J. Visco, L.C. DeJonghe, J. Electrochem. Soc., 135 (1988) 2905-2909.
[140] M. Wu, A. Bhargav, Y. Cui, A. Siegel, M. Agarwal, Y. Ma, Y. Fu, ACS Energy Lett., 1

CR
(2016) 1221-1226.
[141] M. Wu, Y. Cui, A. Bhargav, Y. Losovyj, A. Siegel, M. Agarwal, Y. Ma, Y. Fu, Angew.
Chem. Int. Ed., 55 (2016) 10027-10031.
[142] T. Sotomura, H. Uemachi, K. Takeyama, K. Naoi, N. Oyama, Electrochim. Acta, 37 (1992)
1851-1854.
US
[143] N. Oyama, J.M. Pope, T. Sotomura, J. Electrochem. Soc., 144 (1997) L47-L51.
[144] N. Oyama, T. Tatsuma, T. Sato, T. Sotomura, Nature, 373 (1995) 598-600.
AN
[145] N. Oyama, O. Hatozaki, in: Macromolecular Symposia, Wiley Online Library, 2000, pp.
171-178.
[146] N. Oyama, in: Macromolecular Symposia, Wiley Online Library, 2000, pp. 221-228.
[147] J.E. Park, S. Kim, S. Mihashi, O. Hatozaki, N. Oyama, in: Macromolecular Symposia,
M

Wiley Online Library, 2002, pp. 35-40.


[148] K. Naoi, K.i. Kawase, M. Mori, M. Komiyama, J. Electrochem. Soc., 144 (1997) L173-
L175.
ED

[149] S.-R. Deng, L.-B. Kong, G.-Q. Hu, T. Wu, D. Li, Y.-H. Zhou, Z.-Y. Li, Electrochim. Acta,
51 (2006) 2589-2593.
[150] K. Takechi, Y. Kato, Y. Hase, Adv. Mater . 27 (2015) 2501-2506.
[151] W. Wang, W. Xu, L. Cosimbescu, D. Choi, L. Li, Z. Yang, Chem. Commun., 48 (2012)
PT

6669-6671.
[152] C. Reichardt, T. Welton, Solute-Solvent Interactions, in: C. Reichardt, T. Welton (Eds.)
Solvents and solvent effects in organic chemistry, John Wiley & Sons, Weinheim, 2011, pp. 8-64.
CE

[153] C. Reichardt, T. Welton, Classification of Solvents, in: C. Reichardt, T. Welton (Eds.)


Solvents and solvent effects in organic chemistry, John Wiley & Sons, Weinheim, 2011, pp. 65-
106.
[154] S. Abbott, C.M. Hansen, Hansen solubility parameters in practice, Hansen-Solubility.com,
AC

2008.
[155] C.M. Hansen, Hansen solubility parameters: a user's handbook, CRC press, Boca Raton,
2007.
[156] O.R. Luca, J.L. Gustafson, S.M. Maddox, A.Q. Fenwick, D.C. Smith, Org. Chem. Front., 2
(2015) 823-848.
[157] L. Xing, J. Vatamanu, O. Borodin, G.D. Smith, D. Bedrov, J. Phys. Chem. C, 116 (2012)
23871-23881.
[158] H. Chen, Y.C. Lu, Adv. Energy Mater., 6 (2016) 1502183-1502192.

52
ACCEPTED MANUSCRIPT

[159] K.M. Nouel, P.S. Fedkiw, Electrochim. Acta, 43 (1998) 2381-2387.


[160] J. Xi, Z. Wu, X. Qiu, L. Chen, J. Power Sources, 166 (2007) 531-536.
[161] C. Jia, F. Pan, Y.G. Zhu, Q. Huang, L. Lu, Q. Wang, Sci. Adv., 1 (2015) e1500886.
[162] Z. Yuan, Q. Dai, L. Qiao, Y. Zhao, H. Zhang, X. Li, J. Membrane Sci., 541 (2017) 465-
473.
[163] R. Bones, J. Coetzer, R. Galloway, D. Teagle, J. Electrochem. Soc., 134 (1987) 2379-2382.
[164] S. Stramare, V. Thangadurai, W. Weppner, Chem. Mater., 15 (2003) 3974-3990.
[165] A. Tang, J. Bao, M. Skyllas-Kazacos, J. Power Sources, 196 (2011) 10737-10747.
[166] J.W. Campos, M. Beidaghi, K.B. Hatzell, C.R. Dennison, B. Musci, V. Presser, E.C.

T
Kumbur, Y. Gogotsi, Electrochim. Acta, 98 (2013) 123-130.
[167] H. Chen, Q. Zou, Z. Liang, H. Liu, Q. Li, Y.C. Lu, Nat. Commun., 6 (2015) 5877-5886.

IP
[168] S. Xu, L. Zhang, X.-P. Zhang, Y. Cai, S. Zhang, J. Mater. Chem. A, 5 (2017) 12904-12913.
[169] F.Y. Fan, W.H. Woodford, Z. Li, N. Baram, K.C. Smith, A. Helal, G.H. McKinley, W.C.

CR
Carter, Y.-M. Chiang, Nano Lett., 14 (2014) 2210-2218.
[170] S. Hamelet, D. Larcher, L. Dupont, J.-M. Tarascon, J. Electrochem. Soc., 160 (2013)
A516-A520.
[171] H. Chen, N.-C. Lai, Y.-C. Lu, Chem. Mater., 29 (2017) 7533-7542.

US
[172] S. Hamelet, T. Tzedakis, J.-B. Leriche, S. Sailler, D. Larcher, P.-L. Taberna, P. Simon, J.-
M. Tarascon, J. Electrochem. Soc., 159 (2012) A1360-A1367.
[173] Q. Huang, J. Yang, C.B. Ng, C. Jia, Q. Wang, Energy Environ. Sci., 9 (2016) 917-921.
AN
[174] L. Madec, M. Youssry, M. Cerbelaud, P. Soudan, D. Guyomard, B. Lestriez, J.
Electrochem. Soc., 161 (2014) A693-A699.
[175] Z. Li, K.C. Smith, Y. Dong, N. Baram, F.Y. Fan, J. Xie, P. Limthongkul, W.C. Carter, Y.-
M. Chiang, Phys. Chem. Chem. Phys., 15 (2013) 15833-15839.
M

[176] Q. Huang, H. Li, M. Grätzel, Q. Wang, Phys. Chem. Chem. Phys., 15 (2013) 1793-1797.
[177] Y. Liang, Y. Jing, S. Gheytani, K.-Y. Lee, P. Liu, A. Facchetti, Y. Yao, Nat. Mater., 16
(2017) 841-848.
ED

[178] C. Buhrmester, L. Moshurchak, R.L. Wang, J. Dahn, J. Electrochem. Soc., 153 (2006)
A288-A294.
[179] Z. Chen, K. Amine, Electrochem. Commun., 9 (2007) 703-707.
[180] L. Zhang, Z. Zhang, H. Wu, K. Amine, Energy Environ. Sci., 4 (2011) 2858-2862.
PT

[181] Q. Zou, W. Wang, A. Wang, Z. Yu, K. Yuan, Mater. Lett., 117 (2014) 290-293.
[182] J. Hong, M. Lee, B. Lee, D.-H. Seo, C.B. Park, K. Kang, Nat. Commun., 5 (2014) 5335.
[183] F. Xu, X. Chen, Z. Tang, D. Wu, R. Fu, D. Jiang, Chem. Commun., 50 (2014) 4788-4790.
CE

[184] G.-M. Weng, C.-Y.V. Li, K.-Y. Chan, J. Electrochem. Soc., 160 (2013) A1384-A1389.
[185] L. Suo, Y.-S. Hu, H. Li, M. Armand, L. Chen, Nat. Commun., 4 (2013) ncomms2513.
[186] A. Manthiram, X. Yu, S. Wang, Nat. Rev. Mater., 2 (2017) 16103.
[187] J.E. Bachman, L.A. Curtiss, R.S. Assary, J. Phys. Chem. A, 118 (2014) 8852-8860.
AC

[188] G. Merle, M. Wessling, K. Nijmeijer, J. Membrane. Sci., 377 (2011) 1-35.


[189] X. Li, H. Zhang, Z. Mai, H. Zhang, I. Vankelecom, Energy Environ. Sci., 4 (2011) 1147-
1160.
[190] W. Lau, A. Ismail, N. Misdan, M. Kassim, Desalination, 287 (2012) 190-199.
[191] A.M. Stephan, Eur. Polym. J., 42 (2006) 21-42.
[192] P. Zhao, H. Zhang, H. Zhou, J. Chen, S. Gao, B. Yi, J. Power Sources, 162 (2006) 1416-
1420.
[193] A. Tang, J. McCann, J. Bao, M. Skyllas-Kazacos, J. Power Sources, 242 (2013) 349-356.

53
ACCEPTED MANUSCRIPT

[194] Y. Yan, M. Skyllas-Kazacos, J. Bao, J. Energy Storage, 11 (2017) 104-118.


[195] X. Li, Electrochim. Acta, 170 (2015) 98-109.
[196] Q. Chen, M.R. Gerhardt, M.J. Aziz, J. Electrochem. Soc., 164 (2017) A1126-A1132.
[197] D. Chu, X. Li, S. Zhang, Electrochim. Acta, 190 (2016) 434-445.
[198] Q. Zheng, F. Xing, X. Li, G. Ning, H. Zhang, J. Power Sources, 324 (2016) 402-411.
[199] Q. Xu, T. Zhao, P. Leung, Appl. Energy, 105 (2013) 47-56.
[200] M. Yue, Q. Zheng, H. Zhang, X. Li, X. Ma, AIChE Journal, (2017).
[201] B. Turker, S.A. Klein, E.-M. Hammer, B. Lenz, L. Komsiyska, Energy Convers. Manag.,
66 (2013) 26-32.

T
[202] S. Kim, E. Thomsen, G. Xia, Z. Nie, J. Bao, K. Recknagle, W. Wang, V. Viswanathan, Q.
Luo, X. Wei, J. Power Sources, 237 (2013) 300-309.

IP
CR
Graphical abstract

US
AN
M
ED

With the advantages of low-cost, vast abundance, and high tunability in their structure, redox-
PT

active organic molecules show great advantages in the application of redox flow batteries.
CE
AC

54

Das könnte Ihnen auch gefallen