Paper NT 2022

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

BNL-221586-2021-JAAM

Fluorinated interphase enables reversible aqueous zinc battery chemistries

L. Cao, E. Hu

To be published in "Nature Nanotechnology"

May 2021

Chemistry Department
Brookhaven National Laboratory

U.S. Department of Energy


USDOE Advanced Research Projects Agency - Energy (ARPA-E)

Notice: This manuscript has been authored by employees of Brookhaven Science Associates, LLC under
Contract No. DE-SC0012704 with the U.S. Department of Energy. The publisher by accepting the
manuscript for publication acknowledges that the United States Government retains a non-exclusive, paid-
up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow
others to do so, for United States Government purposes.
DISCLAIMER

This report was prepared as an account of work sponsored by an agency of the


United States Government. Neither the United States Government nor any
agency thereof, nor any of their employees, nor any of their contractors,
subcontractors, or their employees, makes any warranty, express or implied, or
assumes any legal liability or responsibility for the accuracy, completeness, or any
third party’s use or the results of such use of any information, apparatus, product,
or process disclosed, or represents that its use would not infringe privately owned
rights. Reference herein to any specific commercial product, process, or service
by trade name, trademark, manufacturer, or otherwise, does not necessarily
constitute or imply its endorsement, recommendation, or favoring by the United
States Government or any agency thereof or its contractors or subcontractors.
The views and opinions of authors expressed herein do not necessarily state or
reflect those of the United States Government or any agency thereof.
Articles
https://doi.org/10.1038/s41565-021-00905-4

Fluorinated interphase enables reversible aqueous


zinc battery chemistries
Longsheng Cao 1,6, Dan Li1,6, Travis Pollard 2,6, Tao Deng1, Bao Zhang1, Chongyin Yang 1,
Long Chen1, Jenel Vatamanu2, Enyuan Hu 3, Matt J. Hourwitz4, Lin Ma1,2, Michael Ding2, Qin Li1,
Singyuk Hou1, Karen Gaskell1, John T. Fourkas4,5, Xiao-Qing Yang 3, Kang Xu 2 ✉, Oleg Borodin 2 ✉
and Chunsheng Wang 1,4 ✉

Metallic zinc is an ideal anode due to its high theoretical capacity (820 mAh g−1), low redox potential (−0.762 V versus the stan-
dard hydrogen electrode), high abundance and low toxicity. When used in aqueous electrolyte, it also brings intrinsic safety, but
suffers from severe irreversibility. This is best exemplified by low coulombic efficiency, dendrite growth and water consump-
tion. This is thought to be due to severe hydrogen evolution during zinc plating and stripping, hitherto making the in-situ for-
mation of a solid–electrolyte interphase (SEI) impossible. Here, we report an aqueous zinc battery in which a dilute and acidic
aqueous electrolyte with an alkylammonium salt additive assists the formation of a robust, Zn2+-conducting and waterproof
SEI. The presence of this SEI enables excellent performance: dendrite-free zinc plating/stripping at 99.9% coulombic efficiency
in a Ti||Zn asymmetric cell for 1,000 cycles; steady charge–discharge in a Zn||Zn symmetric cell for 6,000 cycles (6,000 h); and
high energy densities (136 Wh kg−1 in a Zn||VOPO4 full battery with 88.7% retention for >6,000 cycles, 325 Wh kg−1 in a Zn||O2
full battery for >300 cycles and 218 Wh kg−1 in a Zn||MnO2 full battery with 88.5% retention for 1,000 cycles) using limited
zinc. The SEI-forming electrolyte also allows the reversible operation of an anode-free pouch cell of Ti||ZnxVOPO4 at 100%
depth of discharge for 100 cycles, thus establishing aqueous zinc batteries as viable cell systems for practical applications.

T
he lithium-ion (Li-ion) battery has become a ubiquitous An effective method to suppress the side reactions on Zn0 is to
commodity in portable electronics, transportation and form in situ a Zn2+-conducting solid–electrolyte interphase (SEI)
grid-storage applications. However, our growing reliance on in non-aqueous electrolytes17,18, and water-in-salt electrolytes
the lithium-ion battery in these industries has exposed its short- (WiSE)3,19. However, WiSE suffer from intrinsic disadvantages (for
comings in safety, energy density and cost driven by limited reserves example, high viscosity and high cost), necessitating the use of
of crucial elements1,2. Zinc metal (Zn0) batteries have emerged as lower concentration aqueous electrolytes to form an interphase on
a promising alternative because of their high volumetric capacity Zn0. Recently, it was reported that a dilute aqueous electrolyte based
(5,855 mAh cm−3 versus 2,061 mAh cm−3 for Li)3,4, non-toxicity, on hydrophobic trifluoromethanesulfonate (OTF−) anions enabled
relatively high abundance and low cost5. Coupling Zn0 with aque- a two-electron oxygen chemistry on the cathode surface by remov-
ous electrolytes offers fast kinetics and high safety compared with ing water from the inner Helmholtz layer20. However, on the anode
non-aqueous counterparts6,7. side, Zn plating/stripping in this dilute aqueous electrolyte still
Most efforts in rechargeable Zn batteries have been dedicated suffers from irreversibility (CE <90%). Since the Zn0 anode is the
to identifying cathode materials serving as hosts for either Zn ion essential component for all Zn0 chemistries including Zn–air bat-
(Zn2+) intercalation8–10 or oxygen/air conversion11,12. However, the teries, improving Zn0 reversibility marks a crucial hurdle for com-
Zn0 electrode suffers from severe irreversibility during repeated mercialization of Zn batteries.
plating/stripping across a broad pH range, as evidenced by low Here, we enhanced Zn plating/stripping CE in aqueous zinc
coulombic efficiency (CE)3,8,9 stemming from the attraction trifluoromethanesulfonate (Zn(OTF)2) electrolyte from 87.6% to
between Zn2+ and water13 even at high ZnCl2 concentrations14,15. 99.9% with trimethylethyl ammonium trifluoromethanesulfonate
Water decomposition forms Zn(OH)42− (zincate) in the local (Me3EtNOTF) as an additive, which engenders in-situ formation of
high pH regions, which subsequently converts to insoluble and a fluorinated and hydrophobic interphase that conducts Zn2+ while
electrochemically inert ZnO (ref. 7). Moreover, the inhomoge- suppressing the hydrogen evolution reaction. This in-situ-formed
neous morphology of Zn(OH)2- and ZnO-based depositions interphase suppresses dendrite growth for up to 6,000 cycles. More
on Zn induces the dendrite growth that compromises cycle importantly, 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF + H2O (where m is
life and safety8,9,16. Large excesses of Zn and electrolyte are molality, mol kg–1) electrolytes also enable the stable cycling of both
required to mitigate anode and water consumption during vanadium phosphate (VOPO4) and O2 cathodes. The Zn||VOPO4
cycling, which decrease the overall energy density and cycle life full cell delivers 136 Wh kg−1 and retains 88.7% of its capacity for
of the batteries. >6,000 cycles, while the Zn||O2 delivers 325 Wh kg−1 for >300 cycles

Department of Chemical and Biomolecular Engineering, University of Maryland, College Park, MD, USA. 2Battery Science Branch, Energy Science Division,
1

Sensor and Electron Devices Directorate, US Army Research Laboratory, Adelphi, MD, USA. 3Chemistry Division, Brookhaven National Laboratory, Upton,
NY, USA. 4Department of Chemistry and Biochemistry, University of Maryland, College Park, MD, USA. 5Institute for Physical Science and Technology,
University of Maryland, College Park, MD, USA. 6These authors contributed equally: Longsheng Cao, Dan Li, Travis Pollard. ✉e-mail: [email protected];
[email protected]; [email protected]

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

b c
6
–64 mV 157 mV
0

Current (mA cm–2)


Current (mA cm–2)
–108 mV
358 mV 4
–0.5 –2
63 µA cm
132 mV
–11 mV 2
–1.0

–1.5 0
–0.2 0 0.2 0.4 1.75 2.00 2.25 2.50 2.75 3.00
Potential (V versus Zn/Zn2+) Potential (V versus Zn/Zn2+)
a 10

0
Current (mA cm )
–2

2.6 V
–0.1 V
–10 Electrochemical window = 2.7 V

4 m Zn(OTF)2 + 0.5 m Me3EtNOTF


–20 4 m Zn(OTF)2

–0.5 0 0.5 1.0 1.5 2.0 2.5 3.0


Potential (V versus Zn/Zn2+)

d e
0.04 0.04
0.5 0.5

1st 2nd 5th 10th 20th 27th


0
0.4 82 mV 50 mV 43 mV 41 mV 39 mV 38 mV 0.4 0
1st 50th 100th 200th 500th 1,000th
113 mV 77 mV 75 mV 58 mV 68 mV 70 mV
–0.04
0.3 0.3 –0.04
Voltage (V)

Voltage (V)

–0.08
0.2 0.2
–0.08
0.18 0.20 0.22
0.1 0.1 0.18 0.20 0.22

0 0

–0.1 –0.1
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Capacity (mAh cm–2) Capacity (mAh cm–2)

f g 0.08
100

100.5
95 0.04
Coulombic efficiency (%)

100.2
90
Potential (V)

99.9
0
85
99.6

80 99.3
800 850 900 950 1,000 –0.04

75 4 m Zn(OTF)2 4 m Zn(OTF)2
4 m Zn(OTF)2 + 0.5 m Me3EtNOTF Short circuit 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF
70 –0.08
0 200 400 600 800 1,000 0 200 400 1,900 2,000 5,900 6,000
Cycle number Test time (h)

Fig. 1 | Electrochemical properties of different electrolytes. a–c, The electrochemical stability window of aqueous electrolytes measured using
polarization scanning at 1 mV s−1 on non-active Ti electrodes between −0.2 V and 3.0 V versus Zn/Zn2+: the overall electrochemical stability window (a)
and magnified views of the regions outlined near anodic (b) and cathodic (c) extremes in a. Purple arrows in b show the hydrogen evolution current at
0.4 V. d,e, Zn-metal plating/stripping profiles on a Ti electrode cycled in 4 m Zn(OTF)2 (d) and 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF (e) with magnified views
of selected cycles in the regions outlined between 0.18 V and 0.22 V at a current density of 0.5 mA cm−2. f, Zn plating/stripping CE with magnified views of
selected cycles between 800th cycle to 1,000th cycle in different electrolytes at 0.5 mA cm−2 and 0.5 mAh cm−2. g, Galvanostatic Zn plating/stripping in a
Zn||Zn symmetric cell at 0.5 mA cm−2 and 0.25 mAh cm−2.

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
a b c

5 µm

50 µm 20 µm 10 µm

d e f

2 µm

50 µm 20 µm 10 µm

g h

nm
64

100 nm 100 nm

Fig. 2 | SEM and TEM imaging of Zn metal after 50 plating/stripping cycles in different electrolytes in a Zn||Zn symmetric cell at 0.5 mA cm−2 and
0.25 mAh cm−2. a–f, SEM images after 50 plating/stripping cycles in 4 m Zn(OTF)2 at scales of 50 μm (a), 20 μm (b) and 10 μm (c) and in
4 m Zn(OTF)2 + 0.5 m Me3EtNOTF at scales of 50 μm (d), 20 μm (e) and 10 μm (f). Insets in a and d show optical images of Zn foils after cycling. In c and
f, red lines indicate the etching depth. g,h, TEM images of the cycled Zn anode surface in 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF (g) and 4 m Zn(OTF)2 (h). In
g, red lines indicate the thickness of the interphase.

since Me3EtNOTF further enhances the reversibility of the two- On the cathode side (Fig. 1c), oxygen evolution was further sup-
electron reaction of the O2 cathodes. Its merits were further evi- pressed by Me3EtNOTF, increasing the oxidation onset potential
denced by the extremely high energy density of 218 Wh kg−1 dem- from 2.55 V to 2.6 V. Overall, a wider stability window of ~2.7 V
onstrated by a Zn||MnO2 full battery, with 88.5% capacity retention was achieved (Fig. 1a) with both cathodic (−0.1 V versus Zn/Zn2+)
achieved after 1,000 cycles using limited Zn. The electrolyte also and anodic (2.6 V versus Zn/Zn2+) limits expanded. The revers-
enables reversible operation of an anode-free Ti||ZnxVOPO4 pouch ibility of the Zn0 electrode in 4 m Zn(OTF)2 electrolytes with and
cell at 100% depth of discharge for 100 cycles. without Me3EtNOTF was compared in Zn||Zn or Zn||Ti cells
(Fig. 1d–g and Supplementary Fig. 2). The addition of Me3EtNOTF
Results and discussion increased the overpotential of Zn plating/stripping on the Ti sur-
Zn plating/stripping reversibility. The electrochemical stabil- face (Fig. 1d,e) due to the increased interphasial resistance and the
ity windows of various aqueous electrolytes at almost identical reduction in ionic conductivity (38.2 mS cm−1) as compared with
pH (Supplementary Table 1) were evaluated on non-active tita- alkylammonium-free electrolyte (46.7 mS cm−1) at room tempera-
nium (Ti) electrodes (Fig. 1a–c), where 0.5 m Me3EtNOTF salt in ture (Supplementary Fig. 3). Even so, the resulting ionic conductiv-
4 m Zn(OTF)2 + H2O electrolytes was found to extend the cathodic ity was still higher than WiSE (12.7 mS cm−1) by a factor of three10.
limit. Closer examination showed that Me3EtNOTF drove the Although Me3EtNOTF slightly decreases the conductivity at room
onset potential for water reduction from 358 mV down to 157 mV temperature, the low-temperature (below −32 °C) conductivity was
(versus Zn/Zn2+) before a minor reduction occurred at −64 mV. improved (Supplementary Fig. 3), and the Zn2+-transference num-
Zn plating occurred at −108 mV (Fig. 1a,b). The hydrogen evolu- ber (t+) was increased to ~0.4 on the basis of Bruce–Vincent meth-
tion current of 63 μA cm−2 at 0.4 V in 4 m Zn(OTF)2 electrolytes odology and 0.32–0.36 from molecular dynamics (MD) simulations
was eliminated with the addition of Me3EtNOTF (purple arrows, (Supplementary Fig. 4, Supplementary note 1 and Supplementary
Fig. 1b). The negatively shifted potential for both Zn deposition and Table 2). Me3EtNOTF also increases thermal stability by 31 °C
hydrogen evolution (Fig. 1a,b) originated from the protection by according to thermogravimetric analysis (Supplementary Fig. 5).
the ammonium cation adsorption and SEI formation from reduc- Most importantly, the presence of 0.5 m Me3EtNOTF promotes SEI
tion of OTF− anions on the anode surface (Supplementary Fig. 1). formation, improving Zn plating/stripping efficiency to >99.0%

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

a b c
F1s F1s F1s
*CF3 ZnF2 ZnF2
(688.6) (684.8) (684.9)

Intensity (a.u.)
Intensity (a.u.)
Intensity (a.u.)

ZnF2
(684.7) *CF3 *CF3
(688.4) (688.0)

692 690 688 686 684 682 692 690 688 686 684 682 692 690 688 686 684 682
Binding energy (eV) Binding energy (eV) Binding energy (eV)

Sputter: 0 min Sputter: 2 min Sputter: 10 min

d e f
C1s C1s C1s
*CF3 C–H
(292.8) C–O C–O
C–C C–O C–H C–N C–H
C–N
(284.8) C–N C–C (286.2) C–C
(286.4)
Intensity (a.u.)

Intensity (a.u.)

(286.4) (284.8)

Intensity (a.u.)
(284.8)
*CF3
ZnCO3
(292.9) *CF
ZnCO3 *CF3 (288.0)
ZnCO3 (289.9)
*CF (288.4) (292.7)
(288.6) (290.1)

294 292 290 288 286 284 282 296 292 288 284 296 292 288 284 280
Binding energy (eV) Binding energy (eV) Binding energy (eV)

Sputter: 0 min Sputter: 2 min Sputter: 10 min

g h i
50 SEI
Zn electrode after 1,000 cycles
40
Atomic concentration (%)

F Zn Electrolyte
ZnS
χ(R ) (Å )
–4

30 Zn O
3
O
ZnF2. 4H2O ZnF2
20

10 Me3EtN+,
ZnC
C O OTF–,
S 3
(R1R2R3)N
Zn metal
0 N
0 2 4 6 8 10 12 0 1 2 3 4
Etching time (min) Poly-species
R (Å)

Fig. 3 | XPS spectra of F1s and C1s for Zn metal after 50 plating/stripping cycles in 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF at a current density of
0.5 mA cm−2. a–f, Depth profiles of F1s generated after Ar+ sputtering for 0 min (a), 2 min (b) and 10 min (c) and of C1s after 0 min (d), 2 min (e) and
10 min (f). g, Composition of SEI after various durations of Ar+ sputtering on Zn metal cycled in 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF. h, Fourier-transformed
extended X-ray absorption fine structure data of ZnF2·4H2O and Zn electrode recovered from 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF after 1,000 cycles at
0.5 mA cm−2 and 0.25 mAh cm−2. Fourier-transformed extended X-ray absorption fine structure for Zn metal is also shown. χ, the normalized oscillatory
part of the absorption coefficient; R, the phase-shift-adjusted bond distance between absorbing atom and its neighbouring atom. i, Cartoon of proposed
Zn2+-conducting SEI, characterized by small nodular particles embedded in a polymeric framework (see Fig. 2g). a.u., arbitrary units.

within the initial 15 cycles, with an average CE of 98.0% at a cur- to 80.9%, failing at the tenth cycle (Supplementary Fig. 6). Note that
rent of 0.5 mA cm−2 and a capacity of 0.5 mAh cm−2. CE further CE increases with increasing currents for identical capacity of Zn
increased to 99.9% after 100 cycles, averaging to 99.8% across 1,000 plating/stripping (Supplementary note 2). Me3EtNOTF rendered
cycles (Fig. 1f) (for a comparison with other reported anodes, see excellent Zn2+ reversibility in Zn||Zn for over 6,000 cycles, while
Supplementary Table 3). By contrast, Zn plating/stripping in 4 m the Me3EtNOTF-free electrolyte failed after 194 cycles (Fig. 1g).
Zn(OTF)2 electrolytes demonstrated a low CE of 87.6%, failing at the The effectiveness of this alkylammonium salt in forming an SEI
20th cycle at 0.5 mA cm−2 and 0.5 mAh cm−2 (Fig. 1f). When areal was also found in another electrolyte (1 m ZnCl2 in water). Without
capacity is increased to 1.0 mAh cm−2 at 0.5 mA cm−2, Zn plating/ Me3EtNOTF, the Zn||Ti cell showed a low CE (~90%) and short
stripping CE in 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF electrolyte still circuit within 52 cycles. With Me3EtNOTF, CE increased to 99.0%
reached >99.6%, while in 4 m Zn(OTF)2 electrolyte, it deteriorated and supported cycling beyond 100 cycles (Supplementary Fig. 7).

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
O –O
-H+
2+
C Zn C O
HO O– –
O
+
Hofmann
CH3 CH3

F F–
elimination
H3C N+ CH2 CH3 + B: N + HB + H 2C CH2 Zn2+
CH3 H3C CH3
O O
– O– O
+
O S S O
F F C
C F C F
O F C C
F F
S O Zn2+ C F Zn2+ F– + O S O– H2 H2
O F O
e– e–
O F
O S C F
O– F
Zn2+

Fig. 4 | Proposed mechanism demonstrating synergistic reactions between triflate and trimethylethyl ammonium to deposit predominantly fluoride
and carbonate-based SEI. Alkylammonium decomposes in the presence of a strong base (B:) such as OH− or CF3−, produced by triflate reduction and
decomposition, to form an amine that is reactive with dissolved CO2 from ambient air to precipitate ZnCO3 SEI (in parentheses). The released ethylene
scavenges reduction products of triflate to precipitate a protective and conducting polyanion species as well as a ZnF2-rich SEI (in parentheses). Reaction
energies and additional details are given in Supplementary Figs. 23–29.

An in-house device (Supplementary Fig. 8a–c) demonstrated a Time-of-flight secondary-ion mass spectrometry additionally con-
rapid decrease in H2 evolution during Zn plating/stripping with firmed the formation of a ZnF2-rich interphase (Supplementary
Me3EtNOTF present (Supplementary Fig. 8d,e). Fig. 13). X-ray absorption spectroscopy (Fig. 3h) revealed that the
interphase is not simply ZnF2, although the exact composition
Zn plating/stripping morphology. Scanning electron microscopy cannot be determined at this time. However, we can clearly iden-
(SEM) and transmission electron microscopy (TEM) were used to tify Zn species therein as Zn2+ rather than Zn0. A ZnF2-rich inter-
investigate the Zn surface after 50 cycles in a Zn||Zn cell containing phase allows Zn2+ to diffuse through while shielding the Zn surface
4 m Zn(OTF)2 electrolyte at 0.5 mA cm−2. Substantial cracking and from water and preventing parasitic reactions. Equally important
dendritic growth were detected on the cycled Zn surface (Fig. 2a,b). is the high ZnF2/Zn interface energy, which suppresses dendrite
The large thickness of deposited Zn (5 µm) is attributed to the highly growth by promoting lateral rather than vertical Zn2+ migration
porous/mossy structure (between the red lines in Fig. 2c), which and deposition5. This is similar to how the LiF/Li interface works21,
was also confirmed by the non-uniform distribution of the elements although the 64 nm interphase observed in the mixed electrolyte is
in energy-dispersive X-ray spectroscopy mapping (Supplementary much thicker than typically observed on graphitic or lithium-metal
Fig. 9). By contrast, Zn deposited from Me3EtNOTF-containing anodes22,23. By contrast, no ZnF2 was found in F1s high-resolution
electrolyte is smooth and compact (Fig. 2d,e), as supported by a spectroscopy of the Zn electrode recovered from the baseline elec-
thin deposited Zn layer (2 µm) in cross-sectional imaging (red trolyte (Supplementary Figs. 14 and 15).
line range in Fig. 2f) and the uniform distribution of elements in In addition to ZnF2, CO32− (~288.6 eV) was also observed in C1s
energy-dispersive X-ray spectroscopy mapping (Supplementary spectra on the Zn electrode recovered from Me3EtNOTF-containing
Fig. 10). As revealed by TEM, a 64 nm ZnF2-rich interphase was electrolyte (Fig. 3d–f and Supplementary Fig. 16), consistent with
observed on the deposited Zn surface when Me3EtNOTF was a previous report24. Trimethylamine evolved from decomposi-
present (Fig. 2g) and resulted in the emergence of semicircles at tion of the alkylammonium assists in this conversion (Fig. 4 and
high frequencies in the electrochemical impedance spectra for Supplementary Fig. 17). A mixture of Zn(OTF)2 and Me3EtNOTF
Me3EtNOTF-containing electrolyte (red line in Supplementary powders was used as a reference for peak assignments of ~292.5 eV
Fig. 11). This SEI serves as an electron barrier preventing the reduc- for CF3, ~286.4 eV for C–O/C–N and ~284.8 eV for C–H/C–C
tion of water while allowing Zn2+ to migrate. No such interphase (Supplementary Fig. 18a). Etching the surface by Ar+ bombard-
was found on the Zn surface recovered from 4 m Zn(OTF)2 electro- ment exposed an underlying interphase consisting of ZnCO3,
lyte after 50 plating/stripping cycles (Fig. 2h). which remained after prolonged (10 min) sputtering (Fig. 3d–f
and Supplementary Fig. 16). Such resistance against sputtering
ZnF2-rich SEI chemistry. The SEI chemistry was investigated could come only from a very dense surface interphase (Fig. 3i),
using X-ray photoelectron spectroscopy (XPS). Inorganic ZnF2 was similar to LiF observed in WiSE19. The presence of ZnCO3 could
detected at ~684.7 eV, and the ratio of inorganic fluorine to organic also contribute to high CE, as CO32− was absent on the Zn sur-
fluorine increased with etching by Ar+ (Fig. 3a–c and Supplementary face recovered from the baseline electrolyte or in the original salts
Fig. 12). The total percentage increased from 22% on the surface (Supplementary Figs. 18 and 19). The same conclusion applies to
to 35% at 2 min and remained relatively steady at 10 min (Fig. 3g). the detected SO32− (Supplementary Figs. 16 and 20), also missing

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

a b
2.0 2.0

1.5 1.5
1st 30th 1st 300th
Voltage (V)

Voltage (V)
0.8 V 1.2 V 0.7 V 1.0 V

1.0 1.0

0.5 1st 5th


1st 5th 0.5
10th 20th 50th 100th
30th 32nd 200th 300th
0 0
0 200 400 600 800 1,000 0 200 400 600 800 1,000
Capacity (mAh g–1) Capacity (mAh g–1)

c d
180 100
0.05 A g−1 0.05 A g−1 120
4 m Zn(OTF)2 + 0.5 m Me3EtNOTF

Coulombic efficiency (%)


0.15 A g−1 90

Capacity (mAh g–1)


150
Capacity (mAh g–1)

−1
0.7 A g 90
80
120 1.5 A g−1
4 m Zn(OTF)2 + 0.5 m Me3EtNOTF
60
2.0 A g−1 70
90
30 4 m Zn(OTF)2
4 m Zn(OTF)2 60
60 4 m Zn(OTF)2 + 0.5 m Me3EtN(OTF)
0 50
0 5 10 15 20 25 30 0 1,000 2,000 3,000 4,000 5,000 6,000
Cycle number Cycle number

e f g
Anode-free cell 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF
160 100
4 m Zn(OTF)2
Specific capacity (mAh g–1)

Coulombic efficiency (%)


80
120
ZnxVOPO4

60
Ti (–) Ti (+) 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF
80
40

40
20

4 m Zn(OTF)2
0 0
Separator and electrolyte
0 20 40 60 80 100
Cycle number

Fig. 5 | Electrochemical performances of Zn–oxygen and Zn-ion batteries. a,b, Cycling performance of the Zn||O2 batteries in 4 m Zn(OTF)2 (a) and
4 m Zn(OTF)2 + 0.5 m Me3EtNOTF (b) with selected cycle number at 50 mA g−1 under 1,000 mA h g−1 (the areal capacity of the cathode was
0.7 mAh cm−2). c,d, Rate performance (c) and cycling stability and CE (d) of the Zn||VOPO4 coin cell in different electrolytes at 2.0 A g−1. e,f, Schematic
(e) and digital picture (f) of the anode-free 50 mAh pouch cell. g, Cycling performance of the anode-free pouch cells at a charge–discharge rate of
0.5 mA cm−2. Specific capacity was calculated based on the electrode mass.

on the Zn electrode recovered from the baseline electrolyte or in the generated. It is expected that any radical species generated from tri-
original salts (Supplementary Figs. 19 and 21). The decreasing peak flate decomposition could be added to ethylene. The polyanion spe-
ratio of Me3EtN+ to electrically neutral (R1R2R3)N (R1, R2 and R3 cies may precipitate on the anode surface and, together with ZnCO3
denote different attached groups in the typical chemistry context) and ZnF2, provide kinetic protection against H2 evolution by repel-
with sputter time indicated the apparent decomposition of Me3EtN+ ling water from Zn0.
(Supplementary Fig. 22). Density functional theory calculations suggest that down-
Initial reduction of triflate was found to favour cleavage of C–F stream reactions resulting from alkylammonium decomposition
or C–S bonds (Supplementary Fig. 23) in accord with previous are responsible for the formation of ZnCO3 (Fig. 4). The alkyl-
reports19,25. Density functional theory calculations show that the ammonium decomposes through Hofmann elimination, with the
defluorination reaction generates a (SO3)F2C. species that reacts base generated as a result of hydrogen evolution reaction (OH−)
with the ethylene in a highly exergonic polymerization reaction or triflate reduction (CF3−) (ref. 26) (Supplementary Figs. 26a
(Supplementary Fig. 24). The reaction shifts the radical to H2C., and 27). A similar mechanism has been observed previously27–29
which is stabilized through adsorption to Zn at the interface or in on Li0 where alkylammonium decomposes through one- or
solution until a second reduction product is added. One possible two-electron reduction and the radical or carbanion is subsequently
product is shown in Supplementary Fig. 25, where ZnF2 is also deprotonated to generate the alkene. The amine readily reacts with

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
a ZnF2-rich SEI
Zn anode

Separator
KFSI additive

Cathode

Electrolyte

b c d 250
*CF3 100
(689.0)
+

Coulombic efficiency (%)


K 200
80

Capacity (mAh g–1)


Unstable
Intensity (a.u.)

O O 150 5.0 C 60
N−
ZnF2
S S (684.6) 100
40
Coulombic efficiency
F F Charge 20
O O 50 Discharge
0
692 690 688 686 684 682
KFSI + Zn + H2O → ZnF2 + R-SO3K 0 200 400 600 800 1,000
Binding energy (eV)
Cycle number

Fig. 6 | Fabrication and electrochemical performances of aritificial ZnF2 SEI. a, Schematic illustration of the in-situ formation of a ZnF2-rich SEI layer
(red dashed lines). b, Chemical structure and reaction of the KFSI. R, nitrogen containing moiety. c, XPS spectra of F1s for Zn metal after plating/stripping in
2 m Zn(OTF)2 + 0.02 m KFSI at a current density of 0.5 mA cm−2. d, Electrochemical performance of the Zn||MnO2 battery enabled by
2 m Zn(OTF)2 + 0.05 m Mn(OTF)2 + 0.02 m KFSI at 5.0 C.

CO2 or HxCO3 (x = 1, 2) in an aqueous environment, leading to the baseline electrolyte (black line in Supplementary Fig. 36a,b). This
formation of CO32− (Supplementary Fig. 17) and is consistent with indicates desorption of alkylammonium ions on Zn electrodes34,35.
other reports30,31. Hydroxides in zincates generated locally near the When the electrode potential is shifted to −20 mV versus Zn/
electrode surface may also decompose the alkylammonium through Zn2+ from open circuit voltage, polarization due to mixed kinetic
a Hofmann elimination (Supplementary Fig. 26b), where the pre- (plating) and diffusion (Zn2+) control was observed. The increased
ferred product is Zn(OH)2(Me3N)2 (Supplementary Fig. 26c). This interfacial impedance is attributed to the SEI. The hydrophobicity
metastable product (Supplementary Fig. 26d) may be the source of of the electrolytes on the Zn surface was also investigated through
trace nitrogen observed in the SEI (Fig. 3g). measurement of the contact angle at open circuit potential and
It was also suggested that anion species may be chemically when polarized (Supplementary Fig. 37a–d). The angle is large for
decomposed through nucleophilic attack by bases such as hydrox- Me3EtNOTF-containing electrolyte and becomes larger once polar-
ide generated at the anode surface32,33. Several mechanisms of ized due to the formation of SEI. The combined effect makes water
OH−/CF3− attacking carbon and/or sulfur in OTF− are explored molecules unwelcome at Zn surfaces.
(Supplementary Figs. 28 and 29). It is observed that all of the reac-
tions considered are exergonic, favouring the formation of sev- Electrochemical performance of Zn-metal full cells. Full
eral products, although mostly with barriers in excess of 2 eV. A Zn||O2 cells were assembled using Zn0 as the anode and carbon
more reasonable barrier of 1.09 eV was found for the formation of black coated carbon paper as the air cathode. This Zn||O2 cell
HCF3 + SO42−. Destabilizing the reactant as described in ref. 33 with has a theoretical energy density of 1,218 Wh kg−1 and has been
OH− generated via hydrogen evolution reaction on the anode sur- widely proposed as a candidate for automotive and grid-storage
face may further decrease the barrier with ZnSOx (x = 3, 4) products applications12,31,36. Although air cathodes could be stabilized by
generated near the anode surface. a fast and reversible two-electron oxygen redox reaction using
The formation of SEI is often correlated with the cation solva- Zn(OTF)2 + H2O electrolyte20, Zn0 reversibility has remained an
tion structure. The 1H nuclear magnetic resonance chemical shifts issue that we address with the Me3EtNOTF-containing electrolyte.
observed in the electrolytes with and without Me3EtNOTF versus Comparing the previously reported Zn||O2 (Supplementary Table 4)
pure water (Supplementary Fig. 30) suggest that 4 m Zn(OTF)2 and Zn||O2 with 4 m Zn(OTF)2 electrolyte (Fig. 5a), Zn||O2 batter-
strongly affects water structure. The effect of Me3EtNOTF is neg- ies with Me3EtNOTF-containing electrolyte and limited Zn anode
ligible. A joint Raman spectroscopy, MD simulation and den- excess (Fig. 5b) display notable cycle life (300 cycles). Cycling
sity functional theory analysis of OTF− bands (Supplementary was conducted under a constant-capacity mode of 1,000 mAh g−1
Figs. 31–35) in 4 m Zn(OTF)2 + Me3EtNOTF showed that Zn2+ (based on the cathode), corresponding to a full-cell energy density
is solvated by six waters with little Zn–OTF contact ion pair for- of 325 Wh kg−1 (based on the cathode and anode). The remark-
mation (Supplementary Fig. 33). The interfacial adsorption/ able depth of discharge37 of 68% for the Zn electrode during the
desorption behaviour of ions on Zn was investigated using electro- Zn||O2 cycling is attributed to the highly reversible Zn plating/
chemical impedance spectroscopy at different potentials. An induc- stripping chemistry. Me3EtNOTF in 4 m Zn(OTF)2 electrolytes
tive arc emerged at low frequency at open circuit potential for Zn0 enable smooth Zn surfaces to be recovered from prolonged cycling
in Me3EtNOTF-containing electrolyte, but was absent on Zn0 in the (Fig. 2d,e), substantially reducing voltage hysteresis in Zn||O2

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

batteries (Fig. 5b) and suppressing H2 evolution potential (Fig. 1b), fluorine ionicity). This pure ZnF2 SEI also dramatically improved
all of which are in sharp contrast to 4 m Zn(OTF)2 electrolytes, the cycling stability of the Zn||MnO2 cell from 38.2% to 95.3% dur-
which display poor performance and high overpotential (Fig. 5a). ing 100 cycles at 0.5 C (Supplementary Fig. 42). Furthermore, the
The decreased polarization in the Zn||O2 full cell and increased Zn||MnO2 cell with an N/P ratio of 2.0 exhibited good long-term
polarization in the observed anode (Supplementary Fig. 11) dem- cycling stability (88.5% after 1,000 cycles at 5.0 C) (Fig. 6d) and
onstrate that Me3EtNOTF also accelerates the oxygen oxidation– rate performance (Supplementary Fig. 43). The high CE of Zn
reduction reactions. Further improvement of the round-trip energy plating/stripping and limited Zn0 enable a high energy density of
efficiency can be achieved by adopting proper bifunctional catalysts 350 Wh kg−1 (218 Wh kg−1 based on the weight of the cathode and
or redox mediators. anode). The in-situ-formed ZnF2 SEI helped to preserve a smooth
To further demonstrate the versatility of Me3EtNOTF, VOPO4 Zn anode surface (Supplementary Fig. 44) and a uniform fluorine
with a theoretical capacity of 165 mAh g−1 was also used as the dispersion (Supplementary Fig. 45).
cathode to couple with the Zn0 electrode. Most previous work has
adopted impractical half-cell configurations using a large Zn0 excess Conclusions
to maintain decent but deceptive cycling stability, leading to sub- A dilute and acidic aqueous Zn electrolyte was formulated with
stantial underutilization of the Zn0 theoretical capacity. When cal- Me3EtNOTF, which enabled highly reversible and dendrite-free
culating cell energy density, the mass of excess Zn0 was excluded, Zn plating/stripping at CE of 99.9% via in-situ formation of
resulting in a high ‘apparent’ energy density (Supplementary Table 5). a composite interphase consisting of ZnF2, ZnCO3, ZnSO3 and
Here, a Ti plate pre-deposited with quantitative Zn0 in the two- polyanions. This Zn2+-conducting SEI suppresses H2 evolu-
fold excess needed for the cathode (negative-to-positive electrode tion and increases the electrochemical stability window of the
capacity (N/P) ratio of 3.0) was chosen as the anode, so that we electrolyte. Leveraging this high reversibility, serval Zn battery
could investigate the effect of Zn loss on cell performance. This chemistries were demonstrated, all of which outperform their
loading represents a rigorous test for electrolyte stability because counterparts using an electrolyte without Me3EtNOTF. The chal-
it requires high Zn-metal utilization and low parasitic reactions lenging Zn||O2 system delivered high reversibility for 300 cycles,
in each cycle. The high potential (1.9 V) and highly catalytic sur- while the Zn||VOPO4 battery delivered excellent cycling stability,
face of VOPO4 present an additional challenge to the anodic sta- retaining ~88.7% of its original capacity during 6,000 cycles at CE
bility of the electrolyte. In Me3EtNOTF-containing aqueous of ~100% despite limited Zn supply. Even in the most challeng-
electrolytes, discharge capacities of 163.1 mAh g−1, 147.1 mAh g−1, ing anode-free configuration, 50 mAh pouch cells demonstrated
125.3 mAh g−1, 102.7 mAh g−1 and 84.9 mAh g−1 were recorded at 80% of capacity retention for 90 cycles. To further prove the effec-
0.05 A g−1, 0.15 A g−1, 0.7 A g−1, 1.5 A g−1 and 2.0 A g−1, respectively tiveness of ZnF2 in improving Zn reversibility, an artificial ZnF2
(Fig. 5c and Supplementary Fig. 38a). When the rate was shifted SEI was deposited on Zn0 and demonstrated in a Zn||MnO2 bat-
back to 0.05 A g−1, the capacity recovered to 163.3 mAh g−1, show- tery using limited Zn0, which displayed an extremely high energy
ing a resilience to the rapid Zn2+ insertion/extraction and stripping/ density of 350 Wh kg−1 and 88.5% capacity retention after 1,000
plating. At 0.05 A g−1, the energy density based on the cathode mass cycles. The identification of this electrolyte and fundamental
is 160 Wh kg−1, and 100 Wh kg−1 with the anode weight included. understanding about the role of alkylammonium opens up an
In the absence of Me3EtNOTF, the rate performance did not expe- avenue to the development of emerging Zn chemistries.
rience an obvious change (Fig. 5c and Supplementary Fig. 38b).
However, the two electrolytes differed in long-term cycling stabil- Online content
ity, with the Me3EtNOTF-containing electrolyte retaining 88.7% of Any methods, additional references, Nature Research report-
the original reversible capacity after 6,000 cycles at 2.0 A g−1 with an ing summaries, source data, extended data, supplementary infor-
average CE of ~99.9% (Fig. 5d) (for a comparison of performance mation, acknowledgements, peer review information; details of
for VOPO4-based Zn-ion batteries, see Supplementary Table 5). author contributions and competing interests; and statements of
By contrast, the cell with 4 m Zn(OTF)2 electrolyte retained <25% data and code availability are available at https://doi.org/10.1038/
capacity within 2,000 cycles (Fig. 5d). Even at small rate (0.15 A g−1), s41565-021-00905-4.
Me3EtNOTF improved capacity retention up to 92.0% after
200 cycles, compared with 48.8% in the baseline (Supplementary Received: 23 December 2020; Accepted: 17 March 2021;
Fig. 39). The most challenging evaluation of the electrolytes was per- Published: xx xx xxxx
formed using an anode-free configuration in a 50 mAh pouch cell
consisting of Ti||ZnxVOPO4 (Fig. 5e–g), where all Zn are pre-stored References
at the cathode side. This cell maintains 80% of its capacity after 1. Turcheniuk, K., Bondarev, D., Singhal, V. & Yushin, G. Ten years left to
redesign lithium-ion batteries. Nature 559, 467–470 (2018).
90 cycles at 0.5 mA cm−2 when Me3EtNOTF was used, demon- 2. Xu, K. Electrolytes and interphases in Li-ion batteries and beyond. Chem.
strating an unprecedented 100% depth-of-discharge performance, Rev. 114, 11503–11618 (2014).
while the cell with baseline electrolyte falls below <80% capacity 3. Wang, F. et al. Highly reversible zinc metal anode for aqueous batteries. Nat.
within 15 cycles. Mater. 17, 543–549 (2018).
4. Parker, J. F. et al. Rechargeable nickel-3D zinc batteries: an energy-dense,
safer alternative to lithium-ion. Science 356, 415–418 (2017).
Artificial ZnF2 SEI. To exclude the influence of ZnCO3 and polyan- 5. Zheng, J. & Archer, L. A. Controlling electrochemical growth of metallic zinc
ions on SEI, we designed a neat ZnF2 SEI. Since bis(fluorosulfonyl) electrodes: toward affordable rechargeable energy storage systems. Sci. Adv. 7,
imide anion (FSI−) decomposes gradually in aqueous solution38, it eabe0219 (2021).
was chosen as the fluorine source (Fig. 6a,b) in an electrolyte satu- 6. Kundu, D. et al. Aqueous vs. nonaqueous Zn-ion batteries: consequences of
the desolvation penalty at the interface. Ener. Env. Sci. 11, 881–892 (2018).
rated with ZnF2. Formation of ZnF2 SEI was identified from XPS 7. Bayer, M. et al. Influence of water content on the surface morphology of zinc
(Fig. 6c). In contrast to short circuits observed for the Zn||Ti and deposited from EMImOTf/water mixtures. J. Electrochem. Soc. 166,
Zn||Zn cells without FSI, remarkable CE of >99.1% was achieved A909–A914 (2019).
in the presence of only 0.02 m potassium bis(fluorosulfonyl) 8. Higashi, S., Lee, S. W., Lee, J. S., Takechi, K. & Cui, Y. Avoiding short circuits
imide (KFSI) in 2 m Zn(OTF)2 (Supplementary Fig. 40a–d). Based from zinc metal dendrites in anode by backside-plating configuration. Nat.
Commun. 7, 11801 (2016).
on depth profiling of the fluorine distribution (Supplementary 9. Zhao, Z. et al. Long-life and deeply rechargeable aqueous Zn anodes enabled
Fig. 41), it is shown that the surface of the Zn0 electrode recovered by a multifunctional brightener-inspired interphase. Energy Environ. Sci. 12,
from the 2 m Zn(OTF)2 + 0.02 m KFSI electrolyte is ZnF2 rich (high 1938–1949 (2019).

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
10. Zhang, L. et al. ZnCl2 ‘water-in-salt’ electrolyte transforms the performance of 27. Kroon, M. C., Buijs, W., Peters, C. J. & Witkamp, G.-J. Decomposition of
vanadium oxide as a Zn battery cathode. Adv. Funct. Mater. 29, 1902653 ionic liquids in electrochemical processing. Green Chem. 8, 241–245 (2006).
(2019). 28. Markevich, E. et al. In situ FTIR study of the decomposition of
11. Luo, M. et al. PdMo bimetallene for oxygen reduction catalysis. Nature 574, N-butyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)amide ionic
81–85 (2019). liquid during cathodic polarization of lithium and graphite electrodes.
12. Fu, J. et al. Electrically rechargeable zinc–air batteries: progress, challenges, Electrochim. Acta 55, 2687–2696 (2010).
and perspectives. Adv. Mater. 29, 1604685 (2017). 29. Preibisch, Y., Horsthemke, F., Winter, M., Nowak, S. & Best, A. S. Is the
13. Chang, N. et al. An aqueous hybrid electrolyte for low-temperature cation innocent? An analytical approach on the cationic decomposition
zinc-based energy storage devices. Energy Environ. Sci. 13, 3527–3535 (2020). behavior of N-butyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)
14. Zhang, C. et al. A ZnCl2 water-in-salt electrolyte for a reversible Zn metal imide in contact with lithium metal. Chem. Mater. https://doi.org/10.1021/acs.
anode. Chem. Commun. 54, 14097–14099 (2018). chemmater.9b04827 (2020).
15. Zhang, Q. et al. Modulating electrolyte structure for ultralow temperature 30. Chowdhury, F. A., Yamada, H., Higashii, T., Goto, k & Onoda, M. CO2
aqueous zinc batteries. Nat. Commun. 11, 4463 (2020). capture by tertiary amine absorbents: a performance comparison study. Ind.
16. Xie, X. et al. Manipulating the ion-transfer kinetics and interface stability for Eng. Chem. Res. 52, 8323–8331 (2013).
high-performance zinc metal anodes. Energy Environ. Sci. 13, 31. Kortunov, P. V., Siskin, M., Paccagnini, M. & Thomann, H. CO2 reaction
503–510 (2020). mechanisms with hindered alkanolamines: control and promotion of reaction
17. Qiu, H. et al. Zinc anode-compatible in-situ solid electrolyte interphase via pathways. Energy Fuels 30, 1223–1236 (2016).
cation solvation modulation. Nat. Commun. 10, 5374 (2019). 32. Yi, Y. et al. Instability at the electrode/electrolyte interface induced
18. Cao, L., Li, D., Deng, T., Li, Q. & Wang, C. Hydrophobic by hard cation chelation and nucleophilic attack. Chem. Mater. 29,
organic-electrolyte-protected zinc anodes for aqueous zinc batteries. Angew. 8504–8512 (2017).
Chem. Int. Ed. 59, 19292–19296 (2020). 33. Nicolas, D. et al. The role of the hydrogen evolution reaction in the
19. Suo, L. et al. ‘Water-in-salt’ electrolyte enables high-voltage aqueous solid-electrolyte interphase formation mechanism for ‘water-in-salt’
lithium-ion chemistries. Science 350, 938–943 (2015). electrolytes. Energy Environ. Sci. 11, 3491–3499 (2018).
20. Sun, W. et al. A rechargeable zinc–air battery based on zinc peroxide 34. Cao, C.-N. On the impedance plane displays for irreversible electrode
chemistry. Science 371, 46–51 (2021). reactions based on the stability conditions of the steady-state. I.
21. Liu, Z. et al. Interfacial study on solid electrolyte interphase at Li metal One state variable besides electrode potential. Electrochim. Acta 35,
anode: implication for Li dendrite growth. J. Electrochem. Soc. 163, 831–836 (1990).
A592–A598 (2016). 35. Zhang, D., Li, L., Cao, L., Yang, N. & Huang, C. Studies of corrosion
22. Nie, M. et al. Role of solution structure in solid electrolyte interphase inhibitors for zinc–manganese batteries: quinoline quaternary ammonium
formation on graphite with LiPF6 in propylene carbonate. J. Phys. Chem. C phenolates. Corros. Sci. 43, 1627–1636 (2001).
117, 25381–25389 (2013). 36. McKubre, M. C. H. & Macdonald, D. D. The dissolution and passivation of
23. Cao, X. et al. Monolithic solid-electrolyte interphases formed in fluorinated zinc in concentrated aqueous hydroxide. J. Electrochem. Soc. 128,
orthoformate-based electrolytes minimize Li depletion and pulverization. Nat. 524–530 (1981).
Energy 4, 796–805 (2019). 37. Parker, J. F., Ko, J. S., Rolison, D. R. & Long, J. W. Translating materials-level
24. Winiarski, J., Tylus, W., Winiarska, K., Szczygieł, I. & Szczygieł, B. XPS and performance into device-relevant metrics for zinc-based batteries. Joule 2,
FT-IR characterization of selected synthetic corrosion products of zinc 2519–2527 (2018).
expected in neutral environment containing chloride ions. J. Spectrosc. 2018, 38. Liu, L. et al. In situ formation of a stable interface in solid-state batteries. ACS
1–14 (2018). Energy Lett. 4, 1650–1657 (2019).
25. Suo, L. et al. ‘Water-in-salt’ electrolyte makes aqueous sodium-ion battery
safe, green, and long-lasting. Adv. Energy Mater. 7, 1701189 (2017).
26. Chen, Y., Cao, Y., Shi, Y., Xue, Z. & Mu, T. Quantitative research on the Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
vaporization and decomposition of [EMIM][Tf2N] by thermogravimetric published maps and institutional affiliations.
analysis–mass spectrometry. Ind. Eng. Chem. Res. 51, 7418–7427 (2012). © The Author(s), under exclusive licence to Springer Nature Limited 2021

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

Methods and CBS-QB3 model chemistries (refs. 42–47). All complexes were immersed in
Materials. Zn(CF3SO3)2 (Zn(OTF)2) was purchased from Tokyo Chemical implicit solvent represented using the polarizable continuum model (water)
Industry. Mn(CF3SO3)2 (Mn(OTF)2) (5% metallic Mn), V2O5 powder (99.99%), solvation model unless otherwise stated48. Raman spectra were computed at the
H3PO4 (85%), acetone (99.9%), KMnO4 (99.0%), MnSO4·H2O (99.0%) and M05-2X/6-31++G(d,p) level.
ethanol were from Sigma-Aldrich. (CH3CH2)(CH3)3N(CF3SO3) (Me3EtNOTF)
was synthesized at the Army Research Laboratory. De-ionized (DI) water (HPLC MD simulations. MD simulations were performed on 4 m Zn(OTF)2 and 4 m
grade) was purchased from Sigma-Aldrich. Zn foil, Ti foil and Ti gauze were Zn(OTF)2 + 0.5 m Me3EtNOTF electrolytes using a simulation cell shown in
purchased from Alfa Aesar. Carbon paper was purchased from Fuel Cell Store. Supplementary Fig. 33a. Composition of the simulation cells and lengths of
VOPO4·2H2O powder was synthesized according to the previous literature39,40. the simulation trajectories are given in Supplementary Table 2 together with
In detail, the mixture of 24 g of V2O5 powder in 140 ml of 85% H3PO4 and 400 ml the predicted transport properties. The double layer of 4 m Zn(OTF)2 + 0.5 m
of DI water was refluxed at 110 °C for 20 h. The VOPO4·2H2O powder with a Me3EtNOTF electrolyte at the model graphite electrode was studied in a separate
yellow-green colour was centrifuged, washed with acetone and kept under ambient set of MD simulations as a function of applied electrode potential using the
conditions. The VOPO4 was obtained by calcining the as-synthesized VOPO4·2H2O set-up shown in Supplementary Fig. 46. The electrodes comprised of three
powder at 500 °C for 6 h in air. The colour of resulting VOPO4 powder was yellow. graphene layers, each layer having 336 carbon atoms and a cross-sectional area of
β-MnO2 nanoparticles were prepared via a hydrothermal method. To be specific, 29.88 Å × 29.57 Å. The distance between electrodes was 85.6 Å. It was chosen such
15.8 g of KMnO4 in 40 ml DI water and 101.4 g of MnSO4·H2O in 20 ml DI water that the electrolyte density in the middle of the simulation cell matched that from
were mixed under stirring for 60 min at room temperature. The mixture was a bulk MD simulation. The equilibration runs were performed at 393 K for 1 ns,
transferred into an 80 ml Teflon-lined autoclave and maintained at 140 °C for 24 h. while the production runs were performed at 363 K for 8 ns.
The product above was centrifuged, then washed thoroughly using water and/or The equations of motion were solved with a time-reversible49 (reference system
ethanol several times before drying at 60 °C for 24 h. propagator algorithms) integrator over the following time resolutions: (1) the
contribution from bonds and angles to the forces were calculated every 0.5 fs,
Material characterizations. The morphologies of the samples were investigated (2) the contribution of dihedrals and non-bonded forces within 7.5 Å cut-off were
by SEM (Hitachi SU-70) equipped with an energy-dispersive X-ray spectroscopy updated every 1.5 fs and (3) the remainder of the forces (reciprocal space Ewald
detector and TEM (JOEL 2100F). All of the samples for ex-situ SEM and TEM and non-bonded forces within 11 Å cut-off for double layer simulations and 16 Å
were recovered from a full aqueous battery in a 2032 coin-cell configuration for bulk simulations) were updated every 3 fs. A Nosé–Hoover50 thermostat was
after electrochemical cycling. Contact angles between the electrolyte and used for temperature control. An archive containing all input files needed to
electrode surface were measured at room temperature using the sessile drop perform bulk MD simulations is included in the Supplementary Information.
method41. Using a contact-angle goniometer (Ramé-Hart, 250-00), a drop The electrolyte was modelled using APPLE&P (ref. 51) polarizable force field
of electrolyte was dispensed from a syringe onto the Zn electrode surface, with induced point dipoles dumped at short distance via the Thole52 model and a
and a drop image was captured. The contact angle was measured between dampening parameter aThole = 0.4. The electrode layer closest to the electrolyte was
the solid/liquid interface and the tangent line to the drop shape at the liquid/ charged at a constant potential using the Siepmann and Reed charge equilibration
vapour interface. Gas chromatography was performed on an Agilent 7890B technique53,54 for electronic conductors. This charge equilibration method
gas chromatograph. 1H nuclear magnetic resonance spectra were acquired on a calculates on the fly the electrode charges such that they minimize the total
Bruker DRX 500 spectrometer at a 1H frequency of 500 MHz, using the chemical electrostatic energy of the system under the constraints of a constant electrostatic
shift of the 1H nucleus in pure water as 4.7 ppm reference. All nuclear magnetic potential imposed on the position of each electrode atom. The fluctuating electrode
resonance measurements were conducted at 298 K. Fourier-transform infrared charges were calculated numerically via self-consistent iterations every 150 fs (50
spectroscopy was performed with a Nicolet 6700 spectrometer and a Golden time steps). The electrode charges were Gaussian-distributed with widths of 0.5 Å
Gate single-reflection monolithic diamond attenuated-total-reflection sample (refs. 54,55). The position of the electrode atoms was constrained during simulations.
cell. Raman spectra were collected with a Horiba Jobin Yvon Labram Aramis
using a 532 nm diode-pumped solid-state laser between 2,600 cm−1 and 100 cm−1. Data availability
Laser power was set at ~150–450 mV, and 400 scans were accumulated with a The data that support the findings of this study are available from the
resolution of 2 cm−1. XPS data were collected using a high-sensitivity XPS (Kratos corresponding authors upon reasonable request.
AXIS 165, Mg Kα radiation). The C1s peak (284.6 eV) was used as the reference
to calibrate other binding energy values. The electrodes were rinsed and dried
under vacuum before characterizations. The contents of different species in the
SEI layers were obtained by fitting the whole XPS spectra using Casa XPS software. References
The depth-reaching rate for the Ar sputtering is estimated to be ~1.0 nm s−1. The 39. Yamamoto, N., Okuhara, T. & Nakato, T. Intercalation compound of
distribution of the constituent elements at different depths in the cycled Zn metal VOPO4·2H2O with acrylamide: preparation and exfoliation. J. Mater. Chem.
was analysed using time-of-flight secondary-ion mass spectroscopy attached to 11, 1858–1863 (2001).
a Ga+ focused-ion beam SEM (Tescan GAIA3) at an accelerated voltage of 20 kV 40. Wang, F. et al. How water accelerates bivalent ion diffusion at the electrolyte/
and 1 nA current. Thermogravimetric analysis experiments were performed with electrode interface. Angew. Chem. Int. Ed. 57, 11978–11981 (2018).
a TA Q600 differential scanning calorimeter. X-ray absorption spectroscopy 41. Horng, P., Brindza, M. R., Walker, R. A. & Fourkas, J. T. Behavior of organic
measurements were performed at the 7-BM beamline of the National Synchrotron liquids at bare and modified silica interfaces. J. Phys. Chem. C 114,
Light Source II at Brookhaven National Laboratory in the fluorescence mode. 394–402 (2010).
42. Frisch, M. J. et al. Gaussian 16, Revision C.01 (Gaussian, Inc., 2016).
Electrochemical measurements. The VOPO4 or MnO2 cathode was prepared 43. Frisch, M. J., Pople, J. A. & Binkley, J. S. Self-consistent molecular orbital
by compressing active material, carbon black and polytetrafluoroethylene at a methods. 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys.
weight ratio of 8/1/1 onto the Ti mesh. For the anode-free pouch cells, the cathode 80, 3265–3269 (1984).
was fabricated by blending VOPO4 powder, Super P carbon and polyvinylidene 44. Zhao, Y., Schultz, N. E. & Truhlar, D. G. Design of density functionals by
fluoride in a weight ratio of 8/1/1 using N-methyl-2-pyrrolidone as solvent. The combining the method of constraint satisfaction with parametrization for
obtained slurry was pasted onto a Ti foil and vacuum dried at 100 °C for 12 h. The thermochemistry, thermochemical kinetics, and noncovalent interactions.
loading mass of active material was ~3 mg cm−2. Filter paper was employed as the J. Chem. Theory Comput. 2, 364–382 (2006).
separator. We used 4 m Zn(OTF)2 and 4 m Zn(OTF)2 + 0.5 m Me3EtNOTF aqueous 45. Zhao, Y. & Truhlar, D. G. The M06 suite of density functionals for main
solutions as electrolytes. Ti plate with pre-deposited Zn served as the anode. group thermochemistry, thermochemical kinetics, noncovalent interactions,
All Zn from the anode was discharged to the cathode before cycling. A higher excited states, and transition elements: two new functionals and systematic
energy density of ~136 Wh kg−1 was achieved. For air batteries, the O2 cathode testing of four M06-class functionals and 12 other functionals. Theor. Chem.
was prepared by mixing 70% carbon black and 30% polyvinylidene difluoride in Acc. 120, 215–241 (2008).
N-methylpyrrolidone, and the slurry mixture was then coated on carbon paper. 46. Chai, J.-D. & Head-Gordon, M. Long-range corrected hybrid density
After coating, the electrodes were dried at 80 °C for 10 min to remove the solvent functionals with damped atom–atom dispersion corrections. Phys. Chem.
before pressing. The electrodes were cut into 1 cm2 sheets, vacuum dried at 100 °C Chem. Phys. 10, 6615–6620 (2008).
for 24 h and weighed before assembly. The mass of zinc and air electrode for the 47. Montgomer, J. A. Jr., Frisch, M. J., Ochterski, J. W. & Petersson, G. A. A
Zn||O2 cell are 1.238 mg and 0.7 mg, respectively. Cyclic voltammetry was carried complete basis set model chemistry. VI. Use of density functional geometries
out using a GAMRY interface 1000E electrochemical work station. The charge– and frequencies. J. Chem. Phys. 110, 2822–2827 (1999).
discharge experiments were performed on a Land BT2000 battery test system at 48. Scalmani, G. & Frisch, M. J. Continuous surface charge polarizable
room temperature. continuum models of solvation. I. General formalism. J. Chem. Phys. 132,
114110 (2010).
Ab initio calculations. Cluster calculations for reaction energies and free energies 49. Martyna, G. J., Tuckerman, M. E., Tobias, D. J. & Klein, M. L. Explicit
were performed with Gaussian 16 rev. C.01 using M05-2X/6-31++G(d,p), M05- reversible integrators for extended systems dynamics. Mol. Phys. 87,
2X/6-311++G(3df,2pd), M06-2X/6-311++G(3df,2pd), wB97XD/6-311++G(3df,2pd) 1117–1157 (1996).

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
50. Hoover, W. G. Canonical dynamics: equilibrium phase-space distributions. User Facility operated for the DOE Office of Science by Brookhaven National Laboratory
Phys. Rev. A 31, 1695–1697 (1985). under contract no. DESC0012704.
51. Borodin, O. Polarizable force field development and molecular dynamics
simulations of ionic liquids. J. Phys. Chem. B 113, 11463–11478 (2009). Author contributions
52. Thole, B. T. Molecular polarizabilities calculated with a modified dipole L. Cao and D.L. designed the experiments and analysed data. K.X. synthesized the
interaction. Chem. Phys. 59, 341–350 (1981). asymmetric ammonium salt. T.P., O.B. and J.V. conducted the calculations. T.D., C.Y.,
53. Siepmann, J. I. & Sprik, M. Influence of surface topology and L. Chen, L.M., Q.L. and S.H. assisted with the material synthesis and characterizations.
electrostatic potential on water/electrode systems. J. Chem. Phys. 102, E.H. and X.-Q.Y. did X-ray absorption spectroscopy measurement and data analysis.
511–524 (1995). M.D. performed conductivity and differential scanning calorimetry measurements. K.G.
54. Reed, S. K., Lanning, O. J. & Madden, P. A. Electrochemical interface assisted with XPS analysis. M.J.H. and J.T.F. assisted with contact-angle testing. K.X., O.B.
between an ionic liquid and a model metallic electrode. J. Chem. Phys. 126, and C.W. conceived and supervised the project. All authors contributed to interpretation
084704 (2007). of the results.
55. Vatamanu, J., Borodin, O. & Smith, G. D. Molecular dynamics simulations of
atomically flat and nanoporous electrodes with a molten salt electrolyte.
Phys. Chem. Chem. Phys. 12, 170–182 (2010). Competing interests
The authors declare no competing interests.
Acknowledgements
C.W. acknowledges funding support from the US Department of Energy (DOE)
through ARPA-E grant DEAR0000389 and the Center of Research on Extreme Batteries. Additional information
Modelling and experimental work at Army Research Laboratory was supported by the Supplementary information The online version contains supplementary material
Joint Center for Energy Storage Research, an Energy Innovation Hub funded by the US available at https://doi.org/10.1038/s41565-021-00905-4.
Department of Energy under cooperative agreement no. W911NF-19-2-0046. E.H. and Correspondence and requests for materials should be addressed to K.X., O.B. or C.W.
X.-Q.Y. are supported by the Assistant Secretary for Energy Efficiency and Renewable
Energy, Vehicle Technology Office of the US DOE through the Advanced Battery Peer review information Nature Nanotechnology thanks the anonymous reviewers for
Materials Research Program under contract no. DE-SC0012704. This research used their contribution to the peer review of this work.
beamline 7-BM of the National Synchrotron Light Source II, a US DOE Office of Science Reprints and permissions information is available at www.nature.com/reprints.

Nature Nanotechnology | www.nature.com/naturenanotechnology

You might also like