nt12dreuar3esd

(Sean Pound) #1

228 | Nature | Vol 579 | 12 March 2020


Article


products was detected (peaks at 6.7, 7.1 and 8.1 ppm; Fig. 4a), suggesting
that the nature and extent of decomposition depends on the reduc-
tion potential. These observations suggest that the decomposition is
an electrochemical reaction with a possible route outlined in Fig. 4c.
A chemical disproportionation reaction with water, as previously
proposed^24 , cannot be excluded; however, the voltage dependence
suggests that the products seen here are formed electrochemically.
By contrast, for DBEAQ, no decomposition products were detected
during a potential hold at 1.4 V and 1.7 V (Extended Data Fig. 9), which
is in agreement both with previous work^13 and with the proposal that
degradation of DBEAQ occurs over long-term cycling via a mecha-
nism involving nucleophilic attack, rather than by degradation of the
reduced species^13.
The in situ NMR technique can be readily used to monitor battery
self-discharge. As shown in Fig. 4b, the open-circuit voltage of a charged
battery (under the protection of flowing N 2 gas) slowly decreases
from 1.3 V to 1.1 V, and then after 17 h a sharp decrease to 0.5 V is seen.
The in situ NMR spectra show that this is caused by the reoxidation of
DHAQ4− to DHAQ3•− and DHAQ2−. The rapid drop in open-circuit volt-
age occurs when the DHAQ2− signals A and B reappear and sharpen,
indicating that the drop is due to complete oxidation of the solu-
tion. To identify the potential oxidant, we developed an in situ mass-
spectrometry technique based on the design of the H cell to moni-
tor the gas evolution (Fig. 4d). A stepped-potential experiment was
performed from 1.2 V to 2.1 V, in increments of 0.1 V. D 2 evolution
commences at 1.2 V, and it is steadily evolved both at higher potentials
and during the rest periods. These two observations suggest that D 2 evo-
lution originates from a chemical reaction, most probably from water
reduction: DHAQ4− + 2D 2 O → DHAQ2− + D 2  + 2OD−. We note that the
redox potential of DHAQ is higher than that of the hydrogen evolu-
tion from water (−0.83 V versus SHE)^13 in a 1 M KOH aqueous solution,
making this reaction thermodynamically unfavourable (unless
substantial pH fluctuations occur). We cannot rule out the further
reaction of the degraded electrolytes, because there is considerable
literature precedent for reactions of AQ-based structures that involve
hydrogen evolution—anthrone dimerization, for example, forms
bianthrone accompanied by hydrogen evolution^25. In addition, some
reported systems are photosensitive, providing another potential
degradation route^26. Further investigation into the various degradation
mechanisms is in progress to understand how to control the stability
of anthrahydroquinones in aqueous media.
In summary, we have demonstrated two in situ NMR metrologies to
study flow batteries. The formation of radicals and fully reduced anions
is directly observed in two AQ-based redox flow battery systems, in
which their equilibrium concentrations are governed by the poten-
tials of the two single-electron-transfer redox processes. The radical
concentrations as a function of SOC were quantified by analysing the
bulk magnetic susceptibility changes, enabling the voltage separation
of the two successive reductions to be extracted. The redox reaction
was found to be coupled with electron transfer between the radicals
and diamagnetic anions, with NMR spectroscopy providing a method
to measure the rates of these reactions. The presence of self-exchange
electron-transfer reactions in organic flow batteries has major implica-
tions because it affects the overall rates of the redox reactions, control-
ling, for example, the comproportionation rate. Electrochemically
triggered decomposition of DHAQ4− to DHA3−/DHAL3− was observed
under specific cycling conditions, but no decomposition of DBEAQ4−
was observed. The real-time observation of reoxidation of DHAQ4− and
hydrogen evolution in these aqueous media indicates that other side
reactions are occurring that involve solvent water and/or degradation
of DHAQ4−. Owing to the simplicity of the on-line NMR setup—which
consists, in essence, of a laboratory-scale redox flow battery and a flow
NMR sampling tube—we expect wide adoption of this technique, which
will help to advance the understanding of various redox chemistries
in flow batteries and other electrochemical systems. Beyond battery


research, we demonstrate a way to study radical species, particularly
at high radical concentrations when hyperfine coupling features in an
EPR spectrum are lost owing to both electron-spin interactions and
electron-transfer reactions. Our work shows that by following spectral
changes in real time, NMR can provide key information concerning
molecular structure, spin-density distributions and intermolecular
electron-hopping rates.

Online content
Any methods, additional references, Nature Research reporting sum-
maries, source data, extended data, supplementary information,
acknowledgements, peer review information; details of author con-
tributions and competing interests; and statements of data and code
availability are available at https://doi.org/10.1038/s41586-020-2081-7.


  1. US Energy Information Administration. Annual Energy Outlook 2018 (US Department of
    Energy, 2018).

  2. Wei, X. et al. Materials and systems for organic redox flow batteries: status and
    challenges. ACS Energy Lett. 2 , 2187–2204 (2017).

  3. Leung, P. et al. Recent developments in organic redox flow batteries: a critical review.
    J. Power Sources 360 , 243–283 (2017).

  4. Yang, B., Hoober-Burkhardt, L., Wang, F., Prakash, G. K. S. & Narayanan, S. R. An
    inexpensive aqueous flow battery for large-scale electrical energy storage based
    on water-soluble organic redox couples. J. Electrochem. Soc. 161 , A1371–A1380
    (2014).

  5. Lin, K. et al. Alkaline quinone flow battery. Science 349 , 1529–1532 (2015).

  6. Tong, L. et al. UV–vis spectrophotometry of quinone flow battery electrolyte for in situ
    monitoring and improved electrochemical modeling of potential and quinhydrone
    formation. Phys. Chem. Chem. Phys. 19 , 31684–31691 (2017).

  7. Lawton, J. S., Aaron, D. S., Tang, Z. & Zawodzinski, T. A. Electron spin resonance
    investigation of the effects of vanadium ions in ion exchange membranes for uses in
    vanadium redox flow batteries. ECS Trans. 41 , 53–56 (2012).

  8. Webster, R. D. In situ electrochemical-NMR spectroscopy. Reduction of aromatic halides.
    Anal. Chem. 76 , 1603–1610 (2004).

  9. Friedl, J. et al. Asymmetric polyoxometalate electrolytes for advanced redox flow
    batteries. Energy Environ. Sci. 11 , 3010–3018 (2018).

  10. Cao, S.-H. et al. In situ monitoring potential-dependent electrochemical process by liquid
    NMR spectroelectrochemical determination: a proof-of-concept study. Anal. Chem. 89 ,
    3810–3813 (2017).

  11. Pecher, O., Carretero-González, J., Griffith, K. J. & Grey, C. P. Materials’ methods: NMR in
    battery research. Chem. Mater. 29 , 213–242 (2017).

  12. Wang, H. et al. In situ NMR spectroscopy of supercapacitors: insight into the charge
    storage mechanism. J. Am. Chem. Soc. 135 , 18968–18980 (2013).

  13. Kwabi, D. G. et al. Alkaline quinone flow battery with long lifetime at pH 12. Joule 2 ,
    1894–1906 (2018); correction 2 , 1907–1908 (2018).

  14. de Boer, E. & MacLean, C. NMR study of electron transfer rates and spin densities in
    p-xylene and p-diethylbenzene anions. J. Chem. Phys. 44 , 1334–1342 (1966); erratum 47 ,
    3102 (1967).

  15. Pell, A. J., Pintacuda, G. & Grey, C. P. Paramagnetic NMR in solution and the solid state.
    Prog. Nucl. Magn. Reson. Spectrosc. 111 , 1–271 (2019).

  16. Evans, D. F. The determination of the paramagnetic susceptibility of substances in
    solution by nuclear magnetic resonance. J. Chem. Soc. 1959 , 2003–2005 (1959).

  17. Johnson, C. S. Jr. Nuclear transverse relaxation in electron-transfer reactions. J. Chem.
    Phys. 39 , 2111–2114 (1963).

  18. de Boer, E. & MacLean, C. Spin densities in the alkyl groups of alkyl-substituted
    naphthalene negative ions, determined by NMR. Mol. Phys. 9 , 191–193 (1965).

  19. Bertini, I., Luchinat, C., Parigi, G. & Ravera, E. NMR of Paramagnetic Molecules 2nd edn
    (Elsevier, 2017).

  20. Rosso, K. M., Smith, D. M. A., Wang, Z., Ainsworth, C. C. & Fredrickson, J. K. Self-exchange
    electron transfer kinetics and reduction potentials for anthraquinone disulfonate. J. Phys.
    Chem. A 108 , 3292–3303 (2004).

  21. Grampp, G., Galán, M. & Sacher, M. Kinetics of photoinduced electron transfer reactions
    of some anthraquinone radical anions with various inorganic ions: comparison with
    Marcus cross-relation. Ber. Bunsenges. Phys. Chem 99 , 111–117 (1995).

  22. Rosokha, S. V. & Kochi, J. K. Continuum of outer- and inner-sphere mechanisms for
    organic electron transfer. Steric modulation of the precursor complex in paramagnetic
    (ion-radical) self-exchanges. J. Am. Chem. Soc. 129 , 3683–3697 (2007).

  23. Meisel, D. & Fessenden, R. W. Electron exchange and electron transfer of semiquinones in
    aqueous solutions. J. Am. Chem. Soc. 98 , 7505–7510 (1976).

  24. Goulet, M.-A. et al. Extending the lifetime of organic flow batteries via redox state
    management. J. Am. Chem. Soc. 141 , 8014–8019 (2019).

  25. Donovan, P. M. & Scott, L. T. 4,11-bisanthenequinone and 10,10′-bianthrone: simple one-
    step syntheses from anthrone. Polycycl. Aromat. Compd. 28 , 128–135 (2008).

  26. Filipescu, N., Avram, E. & Welk, K. D. Photooxidative transformations of anthrone,
    bianthronyl, and bianthrone in acid solution. J. Org. Chem. 42 , 507–512 (1977).
    Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
    published maps and institutional affiliations.
    © The Author(s), under exclusive licence to Springer Nature Limited 2020

Free download pdf