Sie sind auf Seite 1von 8

COMMUNICATION

Redox Flow Batteries www.advmat.de

Phenothiazine-Based Organic Catholyte for High-Capacity


and Long-Life Aqueous Redox Flow Batteries
Changkun Zhang, Zhihui Niu, Sangshan Peng, Yu Ding, Leyuan Zhang, Xuelin Guo,
Yu Zhao,* and Guihua Yu*

redox-active polymers,[4] nitroxide rad-


Redox-active organic materials have been considered as one of the most ical compounds,[5] metal-complexes,[6]
promising “green” candidates for aqueous redox flow batteries (RFBs) due to and heteroaromatic compounds,[7] have
the natural abundance, structural diversity, and high tailorability. However, been explored as suitable materials for
many reported organic molecules are employed in the anode, and molecules with rechargeable solid-electrode batteries and
redox flow batteries (RFBs).
highly reversible capacity for the cathode are limited. Here, a class of hetero­
In aqueous RFBs, a series of the
aromatic phenothiazine derivatives is reported as promising positive materials redox-active organic materials have been
for aqueous RFBs. Among these derivatives, methylene blue (MB) possesses investigated: quiones,[7b,8] viologens,[5,9]
high reversibility with extremely fast redox kinetics (electron-transfer rate alloxazines,[10] phenazines,[11] TEMPOs
constant of 0.32 cm s−1), excellent stability in both neutral and reduced states, (2,2,6,6-tetramethyl-1-piperidinyloxy),[4,5]
and ferrocenes[12] (Figure  1a). Benefited
and high solubility in an acetic-acid–water solvent, leading to a high reversible
from the π-conjugated structure, alloxa-
capacity of ≈71 Ah L−1. Symmetric RFBs based on MB electrolyte demonstrate zine- and phenazine-based nitrogen-
remarkable stability with no capacity decay over 1200 cycles. Even concen- containing heteroaromatic molecules
trated MB catholyte (1.5 m) is still able to deliver stable capacity over hun- exhibit more stable electrochemical revers-
dreds of cycles in a full cell system. The impressive cell performance validates ibility than quione- and viologen-based
the practicability of MB for large-scale electrical energy storage. molecules when served as anolytes.[10a,11]
To date, only TEMPO, ferrocene and their
derivatives display relatively high redox
potential, and have been investigated as
Nowadays, the extensive exploitation of fossil fuels and the catholyte in aqueous RFBs.[4,5,12] However, the low ionic con-
increasing demand for large-scale electrical energy storage ductivity of the ion-selecting membrane in neutral electrolytes
call for sustainable and innovative technologies for generating limits the operating current density and rate performance.
and storing energy from renewable resources.[1] Compared Moreover, the demonstrated RFBs are still largely operated at
with inorganic materials that have been applied in commercial low concentrations though high solubility of these derivatives
batteries, redox-active organic materials are advantageous in have been reported (>2 m). The catholyte beyond 2 m of equiva-
structural diversity, resource sustainability, tunable properties, lent electron concentration has yet to be reported.
and potential material cost.[1b,2] Consequently, various kinds To explore positive redox-active molecules, one can intro-
of organic compounds including carbonyl compounds,[1b,3] duce S atoms or other electron withdrawing groups to increase
the electron affinity of the molecules, and thus to increase the
redox potentials.[3b,13] For instance, when substituting one N of
Dr. C. Zhang, S. Peng, Dr. Y. Ding, L. Zhang, X. Guo, Prof. G. Yu phenazine-based heteroaromatic compounds by S atom to form
Materials Science and Engineering Program thiazine-based compounds, a redox potential nearly 4.0 V was
and Department of Mechanical Engineering achieved in solid-electrode batteries.[7a,14] Adding a second S
The University of Texas at Austin
Austin, TX 78712, USA atom leading to thianthrene-based compounds raises the redox
E-mail: ghyu@austin.utexas.edu potential even higher.[13b,c] As one of the typical thiazine het-
Z. Niu, Prof. Y. Zhao eroaromatic compounds, phenothiazine (PTZ) and its deriva-
Institute of Functional Nano & Soft Materials tives (Figure 1b) are highly bioactive and have widespread use
Jiangsu Key Laboratory for Carbon-Based Functional Materials & Devices in the textile and pharmaceutical industries.[15] Most of the
Collaborative Innovation Centre of Suzhou Nano Science
reported PTZ derivatives possess high redox potential. How-
and Technology
Joint International Research Laboratory of Carbon-Based Functional ever, the application of PTZ-based redox-active molecules is
Materials and Devices mainly studied in nonaqueous systems in which PTZ deriva-
Soochow University tives have limited solubility especially for their charged radical
199 Renai Road, Suzhou Industrial Park, Suzhou, Jiangsu 215123, China states.[16] Besides, nonaqueous RFBs still fall behind the corre-
E-mail: yuzhao@suda.edu.cn
sponding aqueous counterparts in terms of ionic conductivity,
The ORCID identification number(s) for the author(s) of this article
ion-selecting membrane choices, and packaging process.[17] It
can be found under https://doi.org/10.1002/adma.201901052.
is, therefore, highly desirable to develop high capacity aqueous
DOI: 10.1002/adma.201901052 catholyte based on PTZ and its derivatives.

Adv. Mater. 2019, 1901052 1901052  (1 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Figure 1.  a) Comparison of various organic molecules for aqueous RFBs. The left axis represents the equivalent electron concentration (the max-
operated concentration of organic molecules × the number of electrons transferred). DHPS: 7,8-dihydroxyphenazine-2-sulfonic acid.[11] DHBQ:
2,5-dihydroxy-1,4-benzoquinone.[19] DHAQ: 2,6-dihydroxyanthraquinone.[20] 2,6-DBEAQ: 4,4′-((9,10-anthraquinone-2,6-diyl)dioxy)dibutyrate.[21]
ACA: 7/8-carboxylic acid.[10b] FMN-Na: flavin mononucleotide.[10a] MV: methyl viologen.[5] (SPr)2V: 1,1′-bis[3-sulfonatopropyl]-4,4′-bipyridinium.[22]
[(NPr)2TTz]: 4,4′- (thiazolo[5,4-d]thiazole-2,5-diyl)bis(1-(3-(trimethylammonio) propyl)pyridin-1-ium) tetrachloride.[23] [(Me)(NPr)V]Cl3: 1-methyl-10-
[3-(trimethylammonio)propyl]-4,40-bipyridinium trichloride.[9] AQS: anthraquinone-2-sulfonic acid.[24] AQDS: 9,10-anthraquinone-2,7-disulphonic
acid.[7b] BTMAP-Fc: bis((3-trimethylammonio)propyl)ferrocene dichloride.[25] FcNCl: (ferrocenylmethyl)trimethylammonium chloride.[12] HO-TEMPO.[5]
TEMPOSP: TEMPO-4-sulfate potassium salt.[26] TEMPTMA: N,N,N-2,2,6,6-heptamethylpiperidinyl oxy-4-ammonium chloride.[27] b) Molecular struc-
tures of PTZ-based derivatives. c) CV curves of 10 × 10−3 m PTZ-based molecules. MB: methylene blue. CPZ: chlorpromazine. PMZ: promethazine. MB
and Azure A tested at 3.0 m H2SO4 solution. CPZ and PMZ tested at pH = 1 solution.

According to substitution types, PTZ-based derivatives solubility by functionalization of hydrophilic side chains and/or
could be roughly categorized into thiazine dye derivatives optimization of electrolytes. The improved solubility could be
(Type I) and N-substituted phenothiazine derivatives (N-PTZs, achieved via adding an ammonium group at nitrogen or both
Type II) (Figure 1b). The two types exhibit different redox reac- sides of ring (Table S1, Supporting Information). Decreasing
tion mechanisms. Thiazine dye derivatives have a cation in the the symmetry of molecules could benefit to enhance solubility
tricyclic structure connecting one chloride or acetate anion, in water. For instance, when dialkyl groups of MB are substi-
and their redox reactions involve two-electron transfers at the tuted by hydrogens (Azure A), the solubility is improved several
nitrogen position and aromatic rings. Meanwhile for N-PTZs, times greater than that of MB. However, it still cannot meet the
they are able to donate one electron to form a radical-cation at the requirement of RFBs (≈0.6 m in H2SO4 solution), even though
nitrogen position, but further donating another electron would MB was tried for RFBs with unclear battery chemistry and
result in significant structural instability of the molecule.[18] As incompetent cell performance, showing low solubility and poor
shown in Figure 1c and Table S1 in the Supporting Informa- stability of MB in the reported conditions.[7d,28] Moreover, the
tion, the typical thiazine dye derivatives, methylene blue (MB), demonstrated performance lagged far behind of the reported
and Azure A have redox potentials of 0.57 and 0.59 V versus organic aqueous RFBs.[7b,10,11,28] In this work, we found that
NHE (normal hydrogen electrode), respectively, which are MB exhibits a considerably high solubility by rational electro-
higher than the reported quinones and other heteroaromatic lyte optimization (≈1.8 m in H2O + acetic acid (AA) + H2SO4
molecules. The N-PTZs display even higher redox potential mixture and >2.0 m in AA) and no deposition occurred after
(0.86 and 0.92 V versus NHE for chlorpromazine (CPZ) and >100 days storage in 1.5 m MB, and elucidated more clearly
promethazine (PMZ) molecules at pH = 1, respectively). the redox characteristics. Density functional theory (DFT) cal-
Due to a highly hydrophobic tricyclic scaffold, the pristine culation by implicit solvent mode reveals that MB has similar
PTZ is insoluble in water. However, it is possible to enhance its solvation energies in H2O and AA solvents (−41.6 kcal mol−1

Adv. Mater. 2019, 1901052 1901052  (2 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Figure 2.  a) CV curves of 10 × 10−3 m MB molecules at different H2SO4 concentrations. b) Ipa/Ipc ratios, diffusion coefficient (D) and electron-transfer
rate constants (k0) at different H2SO4 concentrations (Ipa and Ipc are the anodic and cathodic currents at 10 mV s−1).

in H2O and −38.6 kcal mol−1 in AA). However, according to of MB and Leuco-MB with two electrons and one proton to bal-
the explicit solvent model that represents more realistic inter- ance the charge. While in relatively strong acidic environment
action between solvent and solute molecules (Figures S2–S3, (pH < 5.6), the redox reactions occur between the configuration
Supporting Information), the calculated adsorption energy of of MB and R-MB with two electrons and two protons to balance
MB in AA solvent (−214 kcal mol−1) is almost 4 times than in the charge with the extra positive charge locating at the N of the
H2O (−56 kcal mol−1). The stronger interaction between AA side chain. At pH = 5.6, there is a balance between the Leuco-
and MB would reduce the lattice energy of MB and increase MB and R-MB.
its solubility in AA. Besides, the strong hydrogen bonding in To better understand the redox mechanism, the nucleus
H2O–H2O molecules also inhibits the binding of H2O with MB independent chemical shift (NICS) and anisotropy of the
molecule, resulting in low solubility in H2O. Further UV–vis induced current density (ACID) calculations were conducted
and Fourier-transform infrared spectroscopy (FTIR) spectra on the electronic structures of different redox products
confirmed that there was no structural change for MB in AA (Figure 3b). ACID is a magnetic indicator of aromaticity for
solvent (Figures S4 andS5, Supporting Information). visualizing ring-currents and electron delocalization.[30] Along
Electrochemical analysis and DFT calculations were per- the periphery of the rings, pristine MB showed clear diamag-
formed to investigate the electrochemical properties of MB netic clockwise ring current, which was composed of one weak
molecules. The redox potential of MB was pH dependent aromatic thiazine ring (NICS(1)zz  =  −6.7 ppm, Figure 3c) and
(Figure  2a). At neutral condition, MB exhibited poor elec- the two aromatic benzenoid rings (NICS(1)zz  =  −16.8 ppm,
trochemical activity and reversibility. With the increasing of Figure 3c). Upon reduction, the benzenoid rings of both Leuco-
acidity by adding H2SO4 (H2SO4 concentration ranging from MB and R-MB exhibited strong aromatic while the central thia-
0.1 to 3.0 m), the redox potential gradually increased from zine ring became antiaromatic, which was in consistent with
0.45 to 0.57 V versus NHE. In addition, the Ipa/Ipc (Ipa and Ipc the ACID plot. Note that the two benzenoid rings of Leuco-
stand for the peak anodic and cathodic currents, respectively) MB and R-MB were more aromatic than MB, suggesting good
was promoted from 0.76 in 0.1 m H2SO4 to 1.02 in 3 m H2SO4 stability of the reduced products. In addition, the Mayer bond
(Figure 2b), revealing significantly improved reversibility. The orders (Figure S10, Supporting Information) indicated that
diffusion coefficient of MB calculated by the Levich equation both Leuco-MB and R-MB exhibited stable rigid molecular
is 2.05 × 10−6 cm2 s−1 in 0.1 m H2SO4. The diffusion coefficient structure though was slightly decreased in bond order for the
had a minor negligible change within the measured range of middle thiazine ring upon reduction.[31]
H2SO4 concentration. Moreover, the electron-transfer rate con- Furthermore, the energy levels of the lowest unoccupied
stant (k0) of MB (Figures S6–S9, Supporting Information) was molecular orbital (LUMO) and the highest occupied molecular
enhanced with the increase of H2SO4 concentration, reaching orbital (HOMO) of MB, R-MB, and Leuco-MB were also cal-
0.32 cm s−1 in 3.0 m H2SO4 which is several orders of magni- culated using DFT. The HOMO plots display that the MB,
tude greater than VO2+/VO2+ redox couple[29] and the reported R-MB, and Leuco-MB are structurally stable because of effec-
redox-active organic materials.[5,7b,10] tive electron delocalization and conjugation (Figure 3d). The
As a heteroaromatic compound, MB molecule exhibits HOMO–HOMO energy level difference between MB and R-MB
two resonance forms, in which the positive charge is located (0.43 eV) is smaller than that of MB and Leuco-MB (0.73 eV),
at the S atom of the middle ring or nitrogen of the side chain suggesting a smaller energy barrier for the redox reactions.[32]
(Figure  3a). The redox process of MB is dependent on the Moreover, R-MB shows a lower energy gap (Eg) than Leuco-MB,
acidity of the electrolyte due to the protonation/deprotonation of indicating that the intrinsic electronic conductivity of R-MB
MB involved in the redox reaction. In a weak acidic environment should be higher than for Leuco-MB. In addition, the chemical
(pH > 5.6), the redox reactions occur between the configuration and electrochemical stability of MB electrolyte was evaluated in

Adv. Mater. 2019, 1901052 1901052  (3 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Figure 3.  a) pH-dependent redox process of MB. b–c) ACID plots and NICS values of MB, R-MB, and Leuco-MB. The molecular planes are placed
perpendicular to the magnetic field vector. Small green arrows are computed current density vectors. The isovalue for the surfaces is 0.05 a.u. The
paratropic/diatropic ring currents indicate antiaromaticity and aromaticity, respectively. The circles labeled by red and blue arrows indicate the dia-
tropic and paratropic ring current flow. d) Optimized structure and calculated energy levels of MB, R-MB, and Leuco-MB using DFT method under
M052x/6-31g(d) package.

acidic solution (0.1 m MB in H2O + AA (volume ratio of 3:1) The symmetric cell based on 1.0 m MB produced a capacity
containing 3.5 m H2SO4). The electrolyte exhibited good sta- of 23.9 Ah L−1 (based on the full electrolytes) with excellent
bility in 2000 CV cycles without obvious decay (Figure S11a, capacity retention of 99.6% after 520 cycles (over 24 days) at
Supporting Information). Even stored in ambient condition for 100 mA cm−2. Both of MB and R-MB molecules had stable
120 days, this electrolyte still showed similar electrochemical molecular structure and there was still no peak shift for the
activity (Figure S11b, Supporting Information), demonstrating UV–vis spectra after resting for above 3 months without any
its excellent chemical stability in acidic condition. inert protection (Figure S13, Supporting Information). The
The symmetric cell is a direct probing method to test elec- good stability of R-MB molecules was also confirmed by the
trochemical stability of MB electrolyte at full charge/discharge unchanged charge/discharge capacity after standing for 48 h in
condition.[33] A symmetric cell containing 0.2 m MB displayed 1.0 m symmetric cell (Figure S14, Supporting Information).
a capacity of 5.32 Ah L−1 at 80 mA cm−2, corresponding to a To evaluate the performance of MB in a full cell, vanadium
utilization of 99.3% (Figure 4a). No capacity decay was observed (II) (denoted as V(II)) anolyte (10 mL, 0.8 m V(II)) was paired
after 1200 cycles at this current density (Figure 4a) as well as with MB catholyte (5 mL, 0.1 m). Excess V(II) anolyte was used
at other current densities (Figure S12, Supporting Informa- to determine the electrochemical stability of catholyte during
tion). The high utilization ratio of MB together with capacity cycling. It is of noting that Lee et al. also assembled a flow cell
retention indicated that MB was highly stable during full oxi- in which MB was taken as anolyte and V(V4+/V5+) as catho-
dation and reduction cycling. Even at high concentration, MB lyte, however the cell performance lagged far behind other
still maintained highly stable capacity with coulombic efficiency reported aqueous systems.[28] In this work, the V-MB flow cell
(CE) and energy efficiency (EE) of nearly 100% (Figure 4b,c). demonstrated capacities of 26.2, 25.6, 25.1, and 23.6 mAh at

Adv. Mater. 2019, 1901052 1901052  (4 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Figure 4.  a) Cycling capacities and CE of the symmetric RFB at 80 mA cm−2 with 0.2 m MB electrolyte (inset: charge/discharge profiles of the symmetric
RFB at different cycles). b) Charge/discharge profiles of the symmetric RFB at different cycles with 1.0 m MB electrolyte at 100 mA cm−2. c) Cycling
capacities and efficiencies of the symmetric RFB with 1.0 m MB electrolyte at 100 mA cm−2. d) Charge–discharge profiles of the V-MB RFB at different
current densities. e) Cycling capacities and efficiencies of the V-MB RFB with 0.1 m MB over 900 cycles at 80 mA cm−2. The data shown are selected
from every tenth of cycle.

60, 80, 100, and 140 mA cm−2, respectively (Figure 4c), and The major reasons are the decreased ionic conductivity of V(II)
had no sign of capacity decay over 900 cycles with CE nearly anolyte (Figure S16, Supporting Information) and the crossover
100% (Figure 4d and Figure S15, Supporting Information). and oxidation of V(II) ions.[34]
The slight fluctuation of capacity should result from the varia- The cell performance in concentrated electrolyte is important
tion of environment temperature. Compared with the capacity when evaluating the potential of a new redox-active material for
retention and CE, the EE gradually decreased with the cycling. practical RFBs. Till now, very limited organic compounds have
The pheno­menon was observed in other organic RFBs.[10b,11] been evaluated with concentration higher than 1 m.[7b,10b,11]

Adv. Mater. 2019, 1901052 1901052  (5 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Figure 5.  a) Charge–discharge profiles of V-MB RFBs at different MB concentrations. b) Cycling efficiencies of V-MB RFB with 1 m MB catholyte at
different current densities (inset: corresponding charge–discharge profiles). c) Cycling capacities and CEs of the V-MB RFBs using 1.2 and 1.5 m MB
catholyte at 80 mA cm−2 (inset: charge–discharge profiles over time of V-MB RFB at 1.5 m MB).

Benefited from high solubility in H2O + AA solvent, the MB inert atmosphere protection and the crossover of vanadium
catholyte was able to deliver capacities of 49, 59, and 70 Ah L−1 ions during long-term operation. Excess anolyte could ensure
in 1.0, 1.2, and 1.5 m MB with the corresponding utilization of the full discharge/charge of the MB catholyte in initial cycling.
91%, 92%, and 87% at 80 mA cm−2, respectively (Figure  5a), However, the oxidation of V(II) and the crossover of vanadium
which are superior to the reported aqueous organic electro- ions resulted in the lack of anolyte as the cycling proceeded,
lytes (Table S2, Supporting Information).[4,7b,10,11] At 1.0 m MB thereby leading to the capacity fading of V-MB RFB. Experi-
catholyte, the EEs of the cell were about 76%, 73%, 70%, and ments confirmed that the capacity could be recovered with
66% at 80, 100, 140, and 200 mA cm−2, respectively, with the refreshed anolyte (Figure S18, Supporting Information).
corresponding CEs of nearly 100% (Figure 5b and Figure S18, Thiazine dye derivatives can be optimized through function-
Supporting Information). Even at 1.5 m, the flow cell delivered alization with electron-withdrawing groups to further increase
performance high catholyte capacity of 71 Ah L−1 at 40 mA cm−2 the redox potential and molecular polarity, thereby achieving
with EE of ≈80% (Figure S17a,b, Supporting Information). high cell voltage and enhanced solubility, which are the two
Cycling tests of the cell with 1.2 and 1.5 m MB catholyte dem- key factors related to the energy density of the proposed cell.
onstrated that the V-MB RFB showed good capacity reten- Selected thiazine dye derivatives were subjected to DFT calcula-
tion after 160 and 50 cycles (with capacity fade of 0.025% per tions on the energy level of the molecular orbitals (Table S3,
cycle and 0.52% per day for 1.2 m, 0.074% per cycle and 0.76% Supporting Information). The LUMOs are linearly related to the
per day for 1.5 m), respectively (Figure 5c,d and Figure S17d, reduction potentials and molecules with more negative LUMO
Supporting Information). The slight fluctuation of capacity and possesses higher redox potential. The energy levels of LUMO
EE was resulted from refreshing of the V(II) anolyte. Compared could be lowered when one N(CH3)2 group in MB molecule
with the long cycling at the symmetric cell (Figure 4b,c), the was substituted by H, NH2, COOH, and N(CH3)(CH2)2NH3+
major reasons for shortened cycling life at high concentra- groups. Moreover, the corresponding solvation energy in water
tion should be attributed to the unstable V(II) anolyte without showed more negative values than MB (Figure S19, Supporting

Adv. Mater. 2019, 1901052 1901052  (6 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Information), suggesting better solubility in water. Further- Rev. 2015, 44, 6684; d) M. R. Lukatskaya, B. Dunn, Y. Gogotsi, Nat.
more, N-PTZ derivatives such as PMZ and CPZ possess high Commun. 2016, 7, 12647.
redox potential and good solubility that could further enhance [2] a) J. Winsberg, T. Hagemann, T. Janoschka, M. D. Hager,
the energy density of RFBs. However, their redox reactions U. S. Schubert, Angew. Chem., Int. Ed. 2017, 56, 686; b) Y. Zhao,
Y. Ding, Y. Li, L. Peng, H. R. Byon, J. B. Goodenough, G. Yu, Chem.
involve the generation of free radicals, which are highly active
Soc. Rev. 2015, 44, 7968.
and may result in decreased stability. Appropriate molecular [3] a) Y. Ding, Y. Li, G. Yu, Chem 2016, 1, 790; b) Q. Zhao, Z. Zhu,
modification and electrode optimization, thus, would be neces- J. Chen, Adv. Mater. 2017, 29, 1607007; c) C. Zhang, Z. Niu, Y. Ding,
sary to tune the chemical properties in the future work.[15a,18] L. Zhang, Y. Zhou, X. Guo, X. Zhang, Y. Zhao, G. Yu, Chem 2018,
In summary, PTZ-based heteroaromatic molecules represent 4, 2814; d) Y. Ding, G. Yu, Angew. Chem., Int. Ed. 2016, 55, 4772;
a promising candidate as redox-active material in aqueous RFBs. e) Y. Ding, G. Yu, Angew. Chem., Int. Ed. 2017, 56, 8614.
As a representative PTZ-based derivative, MB exhibits excep- [4] T. Janoschka, N. Martin, U. Martin, C. Friebe, S. Morgenstern,
tional reversibility as well as structural stability. Electrochemical H. Hiller, M. D. Hager, U. S. Schubert, Nature 2015, 527, 78.
analysis reveals that both MB and its reduced state exhibit excel- [5] T. Liu, X. Wei, Z. Nie, V. Sprenkle, W. Wang, Adv. Energy Mater. 2016,
lent stability and high solubility in acidic solution. When coupled 6, 1501449.
[6] a) Y. Zhao, Y. Ding, J. Song, G. Li, G. Dong, J. B. Goodenough, G. Yu,
with V(II) anolyte, the cell displayed stable capacity retention
Angew. Chem., Int. Ed. 2014, 53, 11036; b) C. Yang, G. Nikiforidis,
in 900 cycles at 0.1 m MB and a high capacity of 71 Ah L−1 J. Y. Park, J. Choi, Y. Luo, L. Zhang, S.-C. Wang, Y.-T. Chan, J. Lim,
based on a 1.5 m MB catholyte. The demonstrated capacity and Z. Hou, M.-H. Baik, Y. Lee, H. R. Byon, Adv. Energy Mater. 2018, 8,
cycling performance are superior to previous organic materials 1702897; c) X. Wei, L. Cosimbescu, W. Xu, J. Z. Hu, M. Vijayakumar,
(Table S2 and Figure S20, Supporting Information). In virtue of J. Feng, M. Y. Hu, X. Deng, J. Xiao, J. Liu, V. Sprenkle, W. Wang, Adv.
the DFT theoretical modeling, better understanding of the PTZ- Energy Mater. 2015, 5, 1400678; d) C. Xie, W. Xu, H. Zhang, X. Hu,
based heteroaromatic class is achieved to help tailor high-per- X. Li, Chem. Commun. 2018, 54, 8419.
formance molecules with high redox potential as well as good [7] a) M. Kolek, F. Otteny, P. Schmidt, C. Muck-Lichtenfeld, C. Einholz,
solubility. The introduction of PTZ-based redox-active molecules J. Becking, E. Schleicher, M. Winter, P. Bieker, B. Esser, Energy
paves the way toward the development of new organic cathode Environ. Sci. 2017, 10, 2334; b) B. Huskinson, M. P. Marshak,
C. Suh, S. Er, M. R. Gerhardt, C. J. Galvin, X. Chen, A. Aspuru-
materials for high-performance aqueous RFBs.
Guzik, R. G. Gordon, M. J. Aziz, Nature 2014, 505, 195; c) C. Zhang,
L. Zhang, Y. Ding, S. Peng, X. Guo, Y. Zhao, G. He, G. Yu, Energy
Storage Mater. 2018, 15, 324; d) A. M. Kosswattaarachchi,
Supporting Information T. R. Cook, ChemElectroChem 2018, 5, 3437.
[8] a) Y. Ding, C. Zhang, L. Zhang, H. Wei, Y. Li, G. Yu, ACS Energy Lett.
Supporting Information is available from the Wiley Online Library or 2018, 3, 2641; b) K. Wedege, E. Dražević, D. Konya, A. Bentien, Sci.
from the author. Rep. 2016, 6, 39101; c) C. Zhang, L. Zhang, Y. Ding, X. Guo, G. Yu,
ACS Energy Lett. 2018, 3, 2875; d) L. Zhang, C. Zhang, Y. Ding,
K. Ramirez-Meyers, G. Yu, Joule 2017, 1, 623.
Acknowledgements [9] C. DeBruler, B. Hu, J. Moss, X. Liu, J. Luo, Y. Sun, T. L. Liu, Chem
2017, 3, 961.
G.Y. acknowledges financial support from Exxon Mobil Corp through its [10] a) A. Orita, M. G. Verde, M. Sakai, Y. S. Meng, Nat. Commun.
membership with University of Texas at Austin, Camille Dreyfus Teacher- 2016, 7, 13230; b) K. Lin, R. Gómez-Bombarelli, E. S. Beh, L. Tong,
Scholar Award, and Sloan Research Fellowship. Y.Z. acknowledges the Q. Chen, A. Valle, A. Aspuru-Guzik, M. J. Aziz, R. G. Gordon, Nat.
financial support from the National Natural Science Foundation of Energy 2016, 1, 16102.
China (Grant No. 51772199), the Collaborative Innovation Centre of [11] A. Hollas, X. Wei, V. Murugesan, Z. Nie, B. Li, D. Reed, J. Liu,
Suzhou Nano Science & Technology, the Base for Introducing Talents V. Sprenkle, W. Wang, Nat. Energy 2018, 3, 508.
of Discipline to Universities, and the Joint International Research [12] B. Hu, C. Debruler, Z. Rhodes, T. Liu, J. Am. Chem. Soc. 2017, 139,
Laboratory of Carbon-Based Functional Materials and Devices.
1207.
[13] a) Y. Ding, C. Zhang, L. Zhang, Y. Zhou, G. Yu, Chem. Soc. Rev.
Conflict of Interest 2018, 47, 69; b) M. E. Speer, M. Kolek, J. J. Jassoy, J. Heine,
M. Winter, P. M. Bieker, B. Esser, Chem. Commun. 2015, 51, 15261;
The authors declare no conflict of interest. c) A. Wild, M. Strumpf, B. Häupler, M. D. Hager, U. S. Schubert,
Adv. Energy Mater. 2017, 7, 1601415; d) Y. Zhao, Y. Ding, J. Song,
L. Peng, J. Goodenough, G. Yu, Energy Environ. Sci. 2014, 7, 1990.
[14] a) G. Dai, X. Wang, Y. Qian, Z. Niu, X. Zhu, J. Ye, Y. Zhao, X. Zhang,
Keywords Energy Storage Mater. 2019, 16, 236; b) C. Peng, G.-H. Ning, J. Su,
catholytes, energy storage, organic redox flow batteries, phenothiazine G. Zhong, W. Tang, B. Tian, C. Su, D. Yu, L. Zu, J. Yang, M.-F. Ng,
Y.-S. Hu, Y. Yang, M. Armand, K. P. Loh, Nat. Energy 2017, 2, 17074.
Received: February 14, 2019 [15] a) S. Mahajan, R. K. Mahajan, Adv. Colloid Interface Sci. 2013,
Revised: April 3, 2019 199–200, 1; b) B. Seger, K. Vinodgopal, P. V. Kamat, Langmuir 2007,
Published online: 23, 5471.
[16] a) J. D. Milshtein, A. P. Kaur, M. D. Casselman, J. A. Kowalski,
S. Modekrutti, P. L. Zhang, N. Harsha Attanayake, C. F. Elliott,
S. R. Parkin, C. Risko, F. R. Brushett, S. A. Odom, Energy Environ.
[1] a) C. P. Grey, J. M. Tarascon, Nat. Mater. 2017, 16, 45; b) Y. Liang, Sci. 2016, 9, 3531; b) H. Chen, Y. Zhou, Y.-C. Lu, ACS Energy
Y. Jing, S. Gheytani, K.-Y. Lee, P. Liu, A. Facchetti, Y. Yao, Nat. Mater. Lett. 2018, 3, 1991; c) A. P. Kaur, N. E. Holubowitch, S. Ergun,
2017, 16, 841; c) Y. Shi, L. Peng, Y. Ding, Y. Zhao, G. Yu, Chem. Soc. C. F. Elliott, S. A. Odom, Energy Technol. 2015, 3, 476; d) J. Huang,

Adv. Mater. 2019, 1901052 1901052  (7 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Z. Yang, M. Vijayakumar, W. Duan, A. Hollas, B. Pan, W. Wang, [24] M. R. Gerhardt, L. Tong, R. Gómez-Bombarelli, Q. Chen,
X. Wei, L. Zhang, Adv. Sustainable Syst. 2018, 2, 1700131; M. P. Marshak, C. J. Galvin, A. Aspuru-Guzik, R. G. Gordon,
e) J. A. Kowalski, M. D. Casselman, A. P. Kaur, J. D. Milshtein, M. J. Aziz, Adv. Energy Mater. 2017, 7, 1601488.
C. F. Elliott, S. Modekrutti, N. H. Attanayake, N. Zhang, [25] E. S. Beh, D. De Porcellinis, R. L. Gracia, K. T. Xia, R. G. Gordon,
S. R. Parkin, C. Risko, F. R. Brushett, S. A. Odom, J. Mater. Chem. M. J. Aziz, ACS Energy Lett. 2017, 2, 639.
A 2017, 5, 24371. [26] J. Winsberg, C. Stolze, A. Schwenke, S. Muench, M. D. Hager,
[17] a) M. Park, J. Ryu, W. Wang, J. Cho, Nat. Rev. Mater. 2017, 2, 16080; U. S. Schubert, ACS Energy Lett. 2017, 2, 411.
b) W. Liu, W. Lu, H. Zhang, X. Li, Chem. - Eur. J. 2019, 25, 1649; [27] T. Janoschka, N. Martin, M. D. Hager, U. S. Schubert, Angew.
c) S. Peng, L. Zhang, C. Zhang, Y. Ding, X. Guo, G. He, G. Yu, Adv. Chem., Int. Ed. 2016, 55, 14427.
Energy Mater. 2018, 8, 1802533; d) L. Zhang, Y. Ding, C. Zhang, [28] W. Lee, Y. Kwon, Korean Chem. Eng. Res. 2018, 56, 890.
Y. Zhou, X. Zhou, Z. Liu, G. Yu, Chem 2018, 4, 1035. [29] H. Fink, J. Friedl, U. Stimming, J. Phys. Chem. C 2016, 120,
[18] M. H. Parvin, Electrochem. Commun. 2011, 13, 366. 15893.
[19] Z. Yang, L. Tong, D. P. Tabor, E. S. Beh, M.-A. Goulet, D. De Porcel [30] a) D. Geuenich, K. Hess, F. Köhler, R. Herges, Chem. Rev. 2005,
linis, A. Aspuru-Guzik, R. G. Gordon, M. J. Aziz, Adv. Energy Mater. 105, 3758; b) Q. Wang, P. Hu, T. Tanaka, T. Y. Gopalakrishna,
2017, 1702056. T. S. Herng, H. Phan, W. Zeng, J. Ding, A. Osuka, C. Chi, J. Siegel,
[20] K. Lin, Q. Chen, M. R. Gerhardt, L. Tong, S. B. Kim, L. Eisenach, J. Wu, Chem. Sci. 2018, 9, 5100.
A. W. Valle, D. Hardee, R. G. Gordon, M. J. Aziz, M. P. Marshak, [31] A. J. Bridgeman, G. Cavigliasso, L. R. Ireland, J. Rothery, J. Chem.
Science 2015, 349, 1529. Soc., Dalton Trans. 2001, 0, 2095.
[21] D. G. Kwabi, K. Lin, Y. Ji, E. F. Kerr, M.-A. Goulet, D. De Porcell [32] T. Ma, Z. Pan, L. Miao, C. Chen, M. Han, Z. Shang, J. Chen, Angew.
inis, D. P. Tabor, D. A. Pollack, A. Aspuru-Guzik, R. G. Gordon, Chem., Int. Ed. 2018, 57, 3158.
M. J. Aziz, Joule 2018, 2, 1894. [33] J. Luo, B. Hu, C. Debruler, Y. Bi, Y. Zhao, B. Yuan, M. Hu, W. Wu,
[22] C. DeBruler, B. Hu, J. Moss, J. Luo, T. L. Liu, ACS Energy Lett. 2018, T. L. Liu, Joule 2019, 3, 149.
3, 663. [34] S. Roe, C. Menictas, M. Skyllas-Kazacos, J. Electrochem. Soc. 2016,
[23] J. Luo, B. Hu, C. Debruler, T. L. Liu, Angew. Chem., Int. Ed. 2018, 57, 231. 163, A5023.

Adv. Mater. 2019, 1901052 1901052  (8 of 8) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Das könnte Ihnen auch gefallen