Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

ARTICLES

PUBLISHED: 18 APRIL 2016 | ARTICLE NUMBER: 16039 | DOI: 10.1038/NENERGY.2016.39

Reversible aqueous zinc/manganese oxide energy


storage from conversion reactions
Huilin Pan1, Yuyan Shao1*, Pengfei Yan2, Yingwen Cheng1, Kee Sung Han2, Zimin Nie1,
Chongmin Wang2, Jihui Yang3, Xiaolin Li1, Priyanka Bhattacharya1, Karl T. Mueller4,5 and Jun Liu1*

Rechargeable aqueous batteries such as alkaline zinc/manganese oxide batteries are highly desirable for large-scale energy
storage owing to their low cost and high safety; however, cycling stability is a major issue for their applications. Here we
demonstrate a highly reversible zinc/manganese oxide system in which optimal mild aqueous ZnSO4 -based solution is used
as the electrolyte, and nanofibres of a manganese oxide phase, α-MnO2 , are used as the cathode. We show that a chemical
conversion reaction mechanism between α-MnO2 and H+ is mainly responsible for the good performance of the system. This
includes an operating voltage of 1.44 V, a capacity of 285 mAh g−1 (MnO2 ), and capacity retention of 92% over 5,000 cycles.
The Zn metal anode also shows high stability. This finding opens new opportunities for the development of low-cost, high-
performance rechargeable aqueous batteries.

E
nergy storage is critical for renewable integration and two-phase reactions have been claimed, for example, spinel
electrification of the energy infrastructure1–8 . Many types ZnMn2 O4 22,25,28 , tunnel-type Znx Mn2 O4 24 , layered birnessite26 ,
of rechargeable battery technologies are being developed. layered Zn-buserite, and/or their complex composite24,29 . Also,
Examples include traditional lead-acid batteries based on Mn K-edge X-ray absorption spectra of Zn/MnO2 show evidence
conversion reactions and Li-ion batteries based on intercalation. only for the redox reaction of Mn during charge/discharge24,26,27 .
Although today’s Li-ion batteries have high energy density, it is still However, the reaction mechanism involved remains a topic
expensive to scale up. On the other hand, lead-acid batteries are of discussion.
low-cost aqueous systems and much easier to scale up for stationary Herein, we demonstrate a highly reversible conversion reaction
applications. Nevertheless, the use of lead causes significant in aqueous Zn/MnO2 systems using α-MnO2 nanofibres as the
environmental concerns besides the low energy density and limited cathode. High capacity and high reversibility can be achieved over
life span, even though lead-acid batteries still account for more than 5,000 cycles, with a capacity retention of 92%. The morphological
half of the global battery market. An alternative, low-cost aqueous and structural evolution of a MnO2 nanofibre cathode were
energy storage system is highly desirable9–11 . carefully investigated during electrochemical reactions by means
So far, a variety of aqueous batteries using alkaline cations, of transmission electron microscopy (TEM), scanning TEM/energy
for example, Li+ , Na+ , K+ , Mg2+ and/or mixed metal ions, as dispersive spectroscopy (STEM–EDS) mapping coupled with X-ray
charge carriers have been reported in the literature12–16 . Recently, diffraction (XRD) methods, and nuclear magnetic resonance
a new concept—aqueous Zn ion batteries—has been studied using (NMR) spectroscopy to reveal the conversion reaction processes.
a high-capacity Zn metal anode (819 mAh g−1 ), whereas most The Zn anode has also been examined carefully. This study provides
previous studies focused on Zn ion intercalation mechanisms and insights necessary to address the challenges faced in the practical use
showed limited long-cycle stability17 . A family of Prussian blue of the Zn/MnO2 system for energy storage.
analogues has lately been reported as zinc intercalation cathodes
in Zn ion batteries. However, these cathodes deliver limited Electrochemical performance of Zn/MnO2 batteries
capacities (around 50 mAh g−1 ) and suffer O2 evolution due to α-MnO2 nanofibres were synthesized with a hydrothermal method
the high operating voltage (about 1.7 V versus Zn)18,19 . Recently, and used as the cathode in a Zn/MnO2 battery chemistry study.
alkaline Zn/MnO2 batteries have been shown to be rechargeable for The XRD pattern in Fig. 1a shows the crystalline phase of α-MnO2
extended cycles using a shallow cycling protocol (typically no more (JCPDS: 44-0141). The TEM image in Fig. 1b shows a homogeneous
than 0.2–0.5 e− per 1 mol MnO2 )20,21 . Meanwhile, fully cycled (1 e− one-dimensional (1D) nanofibre structure of α-MnO2 . The length
per 1 mol MnO2 ) rechargeable Zn/MnO2 cells have been studied of α-MnO2 nanofibre extends to a few micrometres, with a
using α-, γ-phase, or amorphous MnO2 as the cathode in mild diameter of about 50 nm. The inset in Fig. 1b shows high-resolution
aqueous electrolytes22–24 . The MnO2 cathode suffers significant transmission electron microscopy (HRTEM) images, capable of
capacity fading during roughly the initial 20 cycles25–27 . These resolving the lattice, of 1D α-MnO2 nanofibres with a lattice distance
studies also focus on Zn2+ ion intercalation in the MnO2 framework of 0.345 nm for the (220) crystal plane. This indicates that the as-
as the main charge storage mechanism. Several types of intercalation prepared α-MnO2 is highly crystalline, with the [220] axis as the
products through two-phase or a series of complex single- and preferred orientation for 1D fibres30 .

1
Energy & Environment Directorate, Pacific Northwest National Laboratory, Richland, Washington 99352, USA. 2 Environmental Molecular Sciences
Laboratory, Pacific Northwest National Laboratory, Richland, Washington 99352, USA. 3 Department of Materials Science & Engineering, University of
Washington, Seattle, Washington 98195, USA. 4 Physical and Computational Sciences Directorate, Pacific Northwest National Laboratory, Richland,
Washington 99352, USA. 5 Department of Chemistry, Pennsylvania State University, University Park, Pennsylvania 16802, USA.
*e-mail: yuyan.shao@pnnl.gov; jun.liu@pnnl.gov

NATURE ENERGY | VOL 1 | MAY 2016 | www.nature.com/natureenergy 1


© 2016 Macmillan Publishers Limited. All rights reserved
ARTICLES NATURE ENERGY DOI: 10.1038/NENERGY.2016.39

a b

(220)
0.345 nm

10 20 30 40 50 60 70 80
2θ (°) 500 nm

Figure 1 | Structural and morphological characterization of MnO2 . a,b, X-ray diffraction data (a) and TEM image (b) for α-MnO2 nanofibres. The inset of b
shows lattice fringes along the (220) planes.

b 270 270
a

Capacity (mAh g−1)


180 180
1.75

90 90
0 10 20 30
Voltage (V) versus Zn

1.50 0
300 mV 0 100 200 300 400
Cycles
c
60 120

Equivalent capacity
1.25
Zn/Mn mole ratio

loss (mAh g−1)


100
45
First 80
1.00 Second

0 70 140 210 280 30 60


0 5 10 15 20 25
Capacity (mAh g−1)
Cycles

Figure 2 | Electrochemical behaviours of Zn/MnO2 batteries with 2 M ZnSO4 as electrolyte. a,b, Charge/discharge curves of a Zn/MnO2 battery in the
initial two cycles (a) and cycling performance at C/5 (b) of the cell with 2 M ZnSO4 electrolyte (1C = 308 mAh g−1 , capacity is based on the mass of
MnO2 ). The inset in b is an enlargement of the cyclic data during the initial cycles. c, Elemental analysis of dissolved Mn2+ ions in the aqueous electrolyte
during cycling (the red and blue dot lines correspond to the mole ratio between Zn2+ and Mn2+ ions in the electrolyte and the equivalent capacity loss on
cycling, respectively).

In contrast to the conventional primary alkaline MnO2 battery31 , the Mn concentration (measured by the inductively coupled plasma
a mild aqueous ZnSO4 electrolyte was used for the rechargeable (ICP) technique) in the electrolyte solution (Fig. 2c). However, the
Zn/MnO2 battery in this work. Figure 2a shows that α-MnO2 increase in content of dissolved Mn2+ ions slows down after the
nanofibre exhibits an average operating voltage of about 1.44 V initial ten cycles (Fig. 2c). This is probably because the Mn2+ ions
versus Zn and reversible capacities of 210 and 255 mAh g− 1
MnO2
at C/5 gradually dissolved in the electrolyte change the equilibrium of Mn
in the initial two cycles, respectively—higher than the previously dissolution from the MnO2 electrode, and thus suppress significant
reported values22,26 . It is interesting that the Zn/MnO2 battery shows continuous Mn2+ dissolution. This process is consistent with the
rather different charge/discharge curves for the initial two cycles, reduced trend in capacity fading in the following cycles (Fig. 2b).
and a nontrivial overpotential of about 300 mV (C/5) in the initial Finally, a probable self-stabilization in capacity was observed over
cycle, which are rarely seen in intercalation electrode materials32 — hundreds of cycles as a result of the natural dissolution of MnO2 ,
this will be discussed in depth later (note: there is still a overpotential but this takes extended cycling and suffers a significant capacity
gap of about 200 mV even at C/100). The total overpotential is taken loss (Fig. 2b).
roughly as the voltage difference between charge and discharge On the basis of the above observation, we pre-added Mn2+
at the midpoint of the voltage profiles (Supplementary Fig. 1). ions into the electrolyte to change the dissolution equilibrium
Although a high reversible capacity can be delivered in the first of Mn2+ from the MnO2 electrodes, and thus stabilize the
two cycles, a rapid deterioration in capacity was observed for the electrodes themselves. Figure 3a shows cyclic voltammetry (CV)
Zn/MnO2 battery upon cycling (Fig. 2b). We ascribed this capacity curves of Zn/MnO2 batteries in electrolytes, with and without
fading to the dissolution of Mn2+ from Mn3+ disproportionation into MnSO4 , revealing a similar redox behaviour that indicates the
the electrolyte during cycling. This is confirmed by an analysis of MnSO4 additive did not affect the redox reactions in the MnO2

2 NATURE ENERGY | VOL 1 | MAY 2016 | www.nature.com/natureenergy

© 2016 Macmillan Publishers Limited. All rights reserved


NATURE ENERGY DOI: 10.1038/NENERGY.2016.39 ARTICLES
a 800 b 320
No additive C/3
With MnSO4 additive 1C

240
400

Capacity (mAh g−1)


Current (µA)

160 No additive
0.1 M MnSO4 additive
0

80

−400

0
1.00 1.25 1.50 1.75 2.00 0 15 30 45 60
Voltage (V) versus Zn Cycles

c d
100
1.8
100

Coulombic efficiency (%)


Capacity retention (%)
Voltage (V) versus Zn

1.6 5C 92% 75
75

1.4
50
50

1.2

25 25
1.0
10C 5C 2C 1C
0 0
0 70 140 210 280 0 1,000 2,000 3,000 4,000 5,000
Capacity (mAh g−1) Cycles

Figure 3 | Improved electrochemical performance of Zn/MnO2 batteries in optimal aqueous electrolyte. a,b, Comparison of CV scanning (0.1 mV s−1 ,
second cycle) (a) and the cycling performance (b) of MnO2 electrodes with and without 0.1 M MnSO4 additive in a 2 M ZnSO4 aqueous electrolyte at C/3
and 1C, respectively. c,d, Rate performance (c) and the long-term cyclic performance (d) of a Zn/MnO2 battery using an electrolyte with a MnSO4 additive.

electrodes. The two pairs of redox peaks are consistent with the Chemical conversion reaction
two plateaux in the charge/discharge curves shown in Fig. 2a. TEM was used to investigate the morphological and structural
However, with the MnSO4 additive in the electrolyte, the cycling evolution of α-MnO2 electrodes during charge/discharge
performance of the MnO2 electrode is significantly improved in Zn/MnO2 aqueous batteries. Figure 4a shows that the
(Fig. 3b). It is highly likely that pre-addition of 0.1 M Mn2+ α-MnO2 electrode underwent marked changes in morphology
will provide an appropriate equilibrium between Mn2+ dissolution when discharged to 1 V in the first cycle. Pristine well-defined
(from the electrode to the electrolyte, as shown in Fig. 1c) micrometre-long nanofibres were transformed to short nanorods
and the re-oxidation of Mn2+ in the electrolyte, leading to high and nanoparticle aggregates (marked by yellow and blue rectangles
stability of the electrode. Pre-addition of Mn2+ above or below respectively). HRTEM was used to characterize the structural
the equilibrium amount does not address the capacity fading change, as shown in Fig. 4b,c. Short nanorods predominantly
issue well (see detailed dissolution in Supplementary Discussion 1: exhibit a crystalline structure with a lattice distance of 0.33 nm and
Dissolution/oxidation of Mn2+ /MnO2 reactions). Adding MnSO4 to some disordered regions (Fig. 4b). The short nanorods are a newly
the electrolyte also greatly improves the utilization of MnO2 active formed discharge product different from the pristine α-MnO2 , as
material, delivering a higher capacity than that without MnSO4 . can be seen from the difference in the crystalline orientation (arrow
Decent capacities of 285 and 260 mAh g− 1
MnO2
were delivered for indicates the extension outline of the nanorod). In addition, the
cells at current rates of C/3 and 1C, respectively, higher than nanoparticle aggregates in the discharged state consist of ∼5 nm
previously reported values33 . Moreover, the Zn/MnO2 battery with particles with a uniform lattice distance of 0.26 nm, as characterized
the MnSO4 additive shows an excellent rate capability, achieving in Fig. 4c. It was previously reported that Zn2+ ions intercalate
−1
high capacities of 207, 161 and 113 mAh gMnO at 2C, 5C and 10C, 2
into α-MnO2 to form spinel ZnMn2 O4 , or a tunnel or layered
respectively (Fig. 3c). The high rate performance could be ascribed Znx MnO2 phase22,24,26 . However, our TEM results showed that the
to the stabilization and excellent kinetics of the MnO2 electrodes lattice fringes of 0.33 and 0.26 nm from the discharged products
and the Zn anodes (to be discussed below)34–36 . Furthermore, the do not match the reported intercalated phases in the literature.
Zn/MnO2 battery with the MnSO4 additive also exhibits excellent In fact, they are consistent with the d spaces from the (210) and
long-cycle stability, with a high-capacity retention of 92% after (020) planes in monoclinic MnOOH, as evidenced by the XRD
5,000 cycles at a rate of 5C (Fig. 3d). This indicates that a patterns of the MnOOH phase formed shown in Supplementary
Zn/MnO2 system using a mild aqueous electrolyte is very promising Fig. 2. The formation of MnOOH indicates a possible, alternative
for a high-performance, long-life, environmentally-friendly, cost- conversion reaction to Zn2+ ion intercalation into MnO2 . It is
effective energy storage solution. very likely that MnO2 reacts with a proton from water to form

NATURE ENERGY | VOL 1 | MAY 2016 | www.nature.com/natureenergy 3


© 2016 Macmillan Publishers Limited. All rights reserved
ARTICLES NATURE ENERGY DOI: 10.1038/NENERGY.2016.39

a b c d

0.26 nm

0.33 nm
Nanorod 0.26 nm STEM-HAADF

Nanoparticle 0.26 nm
aggregates

0.26 nm
Disordered
200 nm 5 nm zone 5 nm
Mn

e f g

Nanorod

0.34 nm
(220) α-MnO2 O

Nanoparticle
aggregates

100 nm 5 nm 5 nm
Zn

Figure 4 | TEM/HRTEM images of MnO2 electrodes during electrochemical process. a–c,e–g, MnO2 electrodes discharged to 1 V (a–c) and then charged
back to 1.8 V in the first cycle (e–g). The yellow and blue rectangular regions have a morphology typical for short nanorods and nanoparticle aggregates,
respectively. The arrows in b and f indicate the growth directions of the short nanorods. d, STEM-HAADF image of short nanorods and STEM–EDS
mappings of the elemental distributions of Mn, O and Zn in the MnO2 electrode in the discharged state during the first cycle. Electrolyte, 2 M ZnSO4
aqueous electrolyte with 0.1 M MnSO4 additive.

MnOOH (MnO2 + H+ + e− ↔MnOOH). Interestingly, whereas For the charged state, the short nanorods mainly retain their
the H+ ions react with MnO2 , the sequent OH− ions react with morphology, whereas both the lattice distance and crystallinity
ZnSO4 and H2 O in the aqueous electrolyte to form a large flake-like revert to those of the original α-MnO2 electrode, which implies
ZnSO4 [Zn(OH)2 ]3 · xH2 O and reach a neutral charge in the a reversible structure change (Fig. 4d,e). This is consistent with
system. This reaction product was not stable under the high-energy the reversible charge/discharge behaviours of Zn/MnO2 aqueous
electron beam of TEM, but is clearly confirmed in the strong batteries shown above. The nanoparticle aggregations also show
XRD signal (Supplementary Fig. 3a), which dominated the XRD a similar morphology after charging back as they do in the
patterns of other phases, for example, α-MnO2 , and MnOOH, as discharged state. At the same time, the consequent formation of
shown in Supplementary Fig. 2. This reaction was also supported ZnSO4 [Zn(OH)2 ]3 · xH2 O on the cathode side is also reversible
by the 1 H NMR study in Supplementary Fig. 4, which confirms and stable with the hydrogenation/dehydrogenation of the MnO2
the formation of ZnSO4 [Zn(OH)2 ]3 · xH2 O after discharge. The cathode (Supplementary Fig. 3a,d).
above conversion reaction is further confirmed by the STEM–EDS As discussed above, the structure/morphology of α-MnO2
mapping of the MnO2 electrode in the discharged state. Comparing evolves from original micrometre-long nanofibres to a well-mixed
the distribution of the elements Mn, O and Zn in Fig. 4d reveals that composite of short nanorods and nanoparticle aggregations after the
the short nanorods and nanoparticles consist of O and Mn, and no initial cycle, which stabilizes and enhances the structural mechanics
evidence of Zn (note: H atoms cannot be detected in STEM–EDS and kinetics of the electrodes by releasing strain and reducing
mapping). In contrast, Zn is mainly distributed on the flake-like the diffusion length for ions and electrons in the nanocomposite
solid, that is, ZnSO4 [Zn(OH)2 ]3 · xH2 O, as confirmed by the TEM structure, resulting in the excellent rate and cycling stability of
and STEM–EDS mapping in Supplementary Fig. 3b,c. Different the Zn/MnO2 aqueous batteries. In addition, the evolution of the
from previous reports of Zn interaction into MnO2 , both TEM morphology from the original α-MnO2 micro-nanofibres to short
and STEM–EDS mapping evidently confirm the above conversion nanorods and nanoparticle aggregations adds extra surface energy
reaction to form MnOOH. (The background of the Zn mapping to the Gibbs free energy of the α-MnO2 electrode after charging, in
results from associated side products during the formation of comparison to that of the original α-MnO2 fibres; this accounts for
MnOOH on the cathode.) The reaction mechanism of the active the voltage difference between first discharge and second discharge
conversion reaction between MnO2 and H+ is further confirmed as shown in Fig. 2a (see detailed explanation in Supplementary
by the fact that only very limited capacity (capacitor behaviour) Discussion). This phenomenon of a voltage difference in the first
can be delivered in organic Zn2+ -ion-based electrolytes owing to two cycles has also been observed and discussed in other conver-
the absence of H+ ions in organic electrolytes (note: the Zn anode sion reactions from metal oxides electrode materials in lithium-
functions well in an organic electrolyte, as seen from the reversible and sodium-ion batteries37–39 . Furthermore, such thermodynamic
stripping/plating of Zn in Supplementary Fig. 5a). After adding factors, for example, the volume and surface energy variations dur-
H2 O into organic Zn-based electrolytes, the Zn/α-MnO2 battery ing electrochemical processes in the present MnO2 cathode, also
exhibits a similar electrochemical behaviour to that in an aqueous account for the existence of a total overpotential of ∼200 mV in
electrolyte (Supplementary Fig. 5b). the MnO2 electrodes even at a zero-current density (Supplementary

4 NATURE ENERGY | VOL 1 | MAY 2016 | www.nature.com/natureenergy

© 2016 Macmillan Publishers Limited. All rights reserved


NATURE ENERGY DOI: 10.1038/NENERGY.2016.39 ARTICLES
a 40% KOH 2 M ZnSO4 + 0.1 M MnSO4 b

Zn in 40% KOH Zn in ZnSO4 Zn in ZnSO4


12 h 12 h 400+ h
2 Increase in
polarization

5 µm
Voltage (V)

0
0.06
0.24 mA cm−2 0.48 mA cm−2 0.72 mA cm−2 c 342 Zn
0.03 228
−1
Voltage (V)

±38 mV ±35 mV ±35 mV 114


0.00 Zn
O S
0
2 4 6 8 10
−0.03 Energy (keV)
−2

−0.06
0 40 80 120 160 200 240
Time (h)
−3
0 5 10 15 60 120 180 240 5 µm

Test time (h)

Figure 5 | Characterization of Zn anode electrodes. a, Zn stripping/plating from Zn/Zn symmetrical cells at 0.24 mA cm−2 in a 40 wt% alkaline
electrolyte and at 0.24–0.72 mA cm−2 in a mild 2 M ZnSO4 with 0.1 M MnSO4 aqueous electrolyte, respectively. The stripping/plating capacities of the Zn
electrodes (1.25 cm2 ) are 0.3, 0.6 and 0.9 mAh in each cycle. The inset of a shows optical images of cycled Zn anodes in alkaline and mild aqueous
electrolytes. b, SEM image of a pristine Zn anode. c, SEM image and EDX spectrum (inset) of a Zn anode after 120 cycles (240 h) in 2 M ZnSO4 with 0.1 M
MnSO4 aqueous electrolyte.

Fig. 6). This phenomenon is consistent with the significant voltage The reactions of the rechargeable aqueous Zn/MnO2 chemistry
hysteresis of Zn/MnO2 batteries—that is, ∼200 mV of voltage hys- can be formulated as below:
teresis even at C/100 (Supplementary Fig. 1)38 .
Cathode: H2 O ↔ H+ + OH−
MnO2 + H+ + e− ↔ MnOOH
Zn anode 1
Zn2+ + OH− + 16 ZnSO4 + x6 H2 O ↔
The Zn anode was also examined in an effort to understand the 2

high reversibility of the Zn/MnO2 cell chemistry. To a certain


1
6
ZnSO4 [Zn(OH2 )]3 · xH2 O
degree, the reaction of MnO2 + H+ + e− ↔MnOOH in the present Anode: 12 Zn ↔ 12 Zn2+ + e−
Zn/MnO2 battery is similar to that in a MnO2 cathode in a primary Overall: MnO2 + 12 Zn + x6 H2 O + 16 ZnSO4 ↔
alkaline MnO2 battery31 . However, the critical difference lies in the
MnOOH + 16 ZnSO4 [Zn(OH2 )]3 · xH2 O
reaction at the Zn anode side. Figure 5 shows a comparison of
the stability and reversibility of a Zn/Zn symmetric cell in alkaline Based on the total mass of cathode, anode and electrolyte
and mild aqueous electrolytes. In the alkaline electrolyte, there is involved in the above reactions, these reactions provide an energy
a sudden increase in polarization after 12 h for only six cycles of density of ∼170 Wh kg−1 (at C/3). The available energy density
Zn stripping and plating. We ascribe this to the formation of an of the Zn/MnO2 battery is about five times higher than existing
insulating ZnO powder layer to a critical thickness, as shown by commercial lead-acid batteries40 .
the XRD pattern in Supplementary Fig. 7, and a loose and powder- The combination of a highly reversible Zn anode and a MnO2
like of Zn plate surface after cycling in the alkaline electrolyte nanofibre cathode in a mild aqueous electrolyte (that is, 2 M ZnSO4
(inset in Fig. 5a), which results from the irreversible reaction of with 0.1 M MnSO4 in this research) presents a potentially high-
Zn + 2OH− →Zn(OH)2 + 2e− →ZnO + H2 O. In contrast, in capacity aqueous battery chemistry with high reversibility and
a mild 2 M ZnSO4 with 0.1 M MnSO4 aqueous electrolyte, the high cycling stability, that is both environmentally benign and
Zn/Zn symmetric cell shows the excellent kinetics and stability safe. For practical applications, the electrochemical reactions and
of Zn stripping/plating. After more than 120 cycles, the Zn plate stability need to be carefully evaluated. In particular, conversion
still exhibits a smooth, dense surface (inset of Fig. 5a). Figure 5b,c reactions involving both electrode materials and electrolytes are
shows a comparison of SEM images of Zn electrodes before and more complex than intercalation chemistry. The reactions between
after cycling. It can be seen that the Zn electrode exhibits a dense, the cathode, anodes and electrolytes need to be carefully controlled
dendrite-free surface morphology after 120 cycles (Fig. 5b,c). A and optimized. Although this could be a significant challenge, a
similar surface morphology for the Zn anode was also observed good understanding and clarification of the fundamental reactions
in a Zn/MnO2 full cell (Supplementary Fig. 9). This would be in the Zn/MnO2 system is an important step towards a more
very important for the long-term cycling stability of Zn/MnO2 practical system in the future.
batteries. The energy-dispersive X-ray (EDX) spectrum indicates
that the Zn deposition layer surface consists mainly of Zn metal, Conclusion
resulting from the highly reversible plating/stripping of Zn metal in We studied the reversibility of aqueous Zn/MnO2 battery chemistry.
mild electrolyte. The MnO2 nanofibre cathode is highly reversible and stable

NATURE ENERGY | VOL 1 | MAY 2016 | www.nature.com/natureenergy 5


© 2016 Macmillan Publishers Limited. All rights reserved
ARTICLES NATURE ENERGY DOI: 10.1038/NENERGY.2016.39

in a mild ZnSO4 aqueous electrolyte with a MnSO4 additive, 10. Luo, J.-Y., Cui, W.-J., He, P. & Xia, Y.-Y. Raising the cycling stability of aqueous
which suppresses Mn2+ dissolution into the electrolyte. This lithium-ion batteries by eliminating oxygen in the electrolyte. Nature Chem. 2,
Zn/MnO2 cell also exhibits an excellent rate capability and a 760–765 (2010).
11. Lu, Y., Goodenough, J. B. & Kim, Y. Aqueous cathode for next-generation
high-capacity retention of 92% after 5,000 cycles. The structural alkali-ion batteries. J. Am. Chem. Soc. 133, 5756–5759 (2011).
and morphological evolution of α-MnO2 electrodes have been 12. Köhler, J., Makihara, H., Uegaito, H., Inoue, H. & Toki, M. LiV3 O8 :
comprehensively investigated by TEM and STEM–EDS mapping, characterization as anode material for an aqueous rechargeable Li-ion battery
which explains the electrochemical behaviour and reveals an system. Electrochim. Acta 46, 59–65 (2000).
alternative conversion reaction mechanism between MnOOH and 13. Luo, J. Y. & Xia, Y. Y. Aqueous lithium-ion battery LiTi2 (PO4 )3 /LiMn2 O4 with
MnO2 other than Zn2+ ion intercalation into MnO2 . The high high power and energy densities as well as superior cycling stability. Adv. Funct.
Mater. 17, 3877–3884 (2007).
reversibility of the Zn anode was also demonstrated in a mild ZnSO4 14. Wessells, C. D., Huggins, R. A. & Cui, Y. Copper hexacyanoferrate
aqueous electrolyte. The combination of the reversible Zn anode and battery electrodes with long cycle life and high power. Nature Commun.
the MnO2 cathode could lead to a promising battery chemistry that 2, 550 (2011).
is potentially highly reversible, highly stable and safe. 15. Pasta, M., Wessells, C. D., Huggins, R. A. & Cui, Y. A high-rate and long cycle
life aqueous electrolyte battery for grid-scale energy storage. Nature Commun.
3, 1149 (2012).
Methods 16. Chen, L., Zhang, L., Zhou, X. & Liu, Z. Aqueous batteries based on mixed
Material synthesis. In a typical synthesis of α-MnO2 , 0.003 M MnSO4 · H2 O and
monovalence metal ions: a new battery family. ChemSusChem
2 ml 0.5 M H2 SO4 were added to 90 ml deionized water and magnetically stirred
7, 2295–2302 (2014).
until a clear solution was obtained. Then, 20 ml 0.1 M KMnO4 aqueous solution
17. Lee, J.-S. et al. Metal–air batteries with high energy density: Li–air versus
was slowly added into the above solution. The mixture was stirred at room
Zn–air. Adv. Energy Mater. 1, 34–50 (2011).
temperature for 2 h. The solution was then transferred to a Teflon-lined autoclave
18. Zhang, L., Chen, L., Zhou, X. & Liu, Z. Towards high-voltage aqueous
and heated at 120 ◦ C for 12 h. After cooling, the obtained material was collected
metal-ion batteries beyond 1.5 V: the zinc/zinc hexacyanoferrate system. Adv.
by centrifugation, washed three times with water, and dried using a vacuum oven
Energy Mater. http://dx.doi.org/10.1002/aenm.201400930 (2015).
at room temperature.
19. Trócoli, R. & La Mantia, F. An aqueous zinc-ion battery based on copper
hexacyanoferrate. ChemSusChem 8, 481–485 (2015).
Characterizations. X-ray diffraction (XRD) measurements were performed using
20. Mondoloni, C., Laborde, M., Rioux, J., Andoni, E. & Lévy-Clément, C.
a Rigaku Miniflex II diffractometer with Cu Kα radiation (λ = 1.5406 Å). An FEI
Rechargeable alkaline manganese dioxide batteries: I. In situ X-ray diffraction
Titan 80–300 microscope equipped with an objective lens corrector at 300 kV was
investigation of the (EMD-type) insertion system. J. Electrochem. Soc. 139,
employed for the TEM images. Scanning/TEM and EDS mapping were
954–959 (1992).
performed on a JEOL ARM200CF microscope at 200 kV, which is equipped with
21. Hertzberg, B., Sviridov, L., Stach, E. A., Gupta, T. & Steingart, D.
a probe corrector and high-efficiency silicon drift EDS detector with a 100 mm2
A manganese-doped barium carbonate cathode for alkaline batteries.
X-ray sensor. 1 H Nuclear magnetic resonance (NMR) analyses were conducted
J. Electrochem. Soc. 161, A835–A840 (2014).
using a 1.6 mm HX probe using single-pulse excitation on a 600 MHz NMR
22. Xu, C., Li, B., Du, H. & Kang, F. Energetic zinc ion chemistry: the rechargeable
spectrometer (Agilent) under magic angle spinning at 35 kHz.
zinc ion battery. Angew. Chem. 124, 957–959 (2012).
23. Xu, C., Du, H., Li, B., Kang, F. & Zeng, Y. Reversible insertion properties of zinc
Electrochemical measurements. Using an electrochemical process, α-MnO2
ion into manganese dioxide and its application for energy storage. Electrochem.
electrodes were prepared by mixing α-MnO2 powder, carbon black and poly
Solid State Lett. 12, A61–A65 (2009).
vinylidene fluoride at a weight ratio of 7:2:1 in N -methyl pyrrolidone solvent,
24. Alfaruqi, M. H. et al. Electrochemically induced structural transformation in a
then coating the slurry onto carbon sheets. The α-MnO2 mass loading is
γ-MnO2 cathode of a high capacity zinc-ion battery system. Chem. Mater. 27,
1–5 mg cm−2 (Supplementary Fig. 8). Zn/MnO2 batteries were assembled using
3609–3620 (2015).
an α-MnO2 electrode as the cathode, glass fibre as the separator and Zn foil as
25. Xu, D. et al. Preparation and characterization of MnO2 /acid-treated CNT
the anode in CR2032 coin cells. The electrolyte used was 2 M ZnSO4
nanocomposites for energy storage with zinc ions. Electrochim. Acta 133,
with/without 0.1 M MnSO4 as an additive in H2 O. The conventional alkaline
254–261 (2014).
MnO2 batteries were assembled using a similar process, except for the use of
26. Lee, B. et al. Electrochemically-induced reversible transition from the tunneled
40 wt% KOH aqueous electrolyte. A symmetrical Zn/Zn cell was assembled using
to layered polymorphs of manganese dioxide. Sci. Rep. 4, 6066 (2014).
Zn foil as both working and counter electrodes, using 2 M ZnSO4 with 0.1 M
27. Alfaruqi, M. H. et al. Enhanced reversible divalent zinc storage in a structurally
MnSO4 as the electrolyte. 0.1 M ZnTf2 was dissolved into dimethyl sulfoxide
stable α-MnO2 nanorod electrode. J. Power Sources 288, 320–327 (2015).
(DMSO) as an organic Zn electrolyte. For the water-addition organic electrolyte,
28. Xu, C., Chiang, S. W., Ma, J. & Kang, F. Investigation on zinc ion storage in
5 M H2 O was added into the above electrolyte. The charge/discharge was carried
alpha manganese dioxide for zinc ion battery by electrochemical impedance
out on a LANHER battery tester (Wuhan).
spectrum. J. Electrochem. Soc. 160, A93–A97 (2013).
29. Lee, B. et al. Elucidating the intercalation mechanism of zinc ions into
Received 28 October 2015; accepted 4 March 2016; [α]-MnO2 for rechargeable zinc batteries. Chem. Commun.
published 18 April 2016 51, 9265–9268 (2015).
30. Hongen, W., Zhouguang, L., Dong, Q., Yujie, L. & Wei, Z. Single-crystal
α-MnO2 nanorods: synthesis and electrochemical properties. Nanotechnology
References 18, 115616 (2007).
1. Tarascon, J. M. & Armand, M. Issues and challenges facing rechargeable 31. Cheng, F. Y., Chen, J., Gou, X. L. & Shen, P. W. High-power alkaline Zn–MnO2
lithium batteries. Nature 414, 359–367 (2001). batteries using γ-MnO2 nanowires/nanotubes and electrolytic zinc powder.
2. Armand, M. & Tarascon, J. M. Building better batteries. Nature Adv. Mater. 17, 2753–2756 (2005).
451, 652–657 (2008). 32. Sun, J. et al. Overpotential and electrochemical impedance analysis on Cr2 O3
3. Liu, J. et al. Materials science and materials chemistry for large scale thin film and powder electrode in rechargeable lithium batteries. Solid State
electrochemical energy storage: from transportation to electrical grid. Adv. Ion. 179, 2390–2395 (2008).
Funct. Mater. 23, 929–946 (2013). 33. Alias, N. & Mohamad, A. A. Advances of aqueous rechargeable lithium-ion
4. Larcher, D. & Tarascon, J. M. Towards greener and more sustainable batteries battery: a review. J. Power Sources 274, 237–251 (2015).
for electrical energy storage. Nature Chem. 7, 19–29 (2015). 34. Hu, Y. S., Kienle, L., Guo, Y. G. & Maier, J. High lithium electroactivity of
5. Jiang, J. et al. Recent advances in metal oxide-based electrode architecture nanometer-sized rutile TiO2 . Adv. Mater. 18, 1421–1426 (2006).
design for electrochemical energy storage. Adv. Mater. 24, 5166–5180 (2012). 35. Chan, C. K. et al. High-performance lithium battery anodes using silicon
6. Qu, D. Studies of the activated carbons used in double-layer supercapacitors. nanowires. Nature Nanotech. 3, 31–35 (2008).
J. Power Sources 109, 403–411 (2002). 36. Lee, H.-W. et al. Ultrathin spinel LiMn2 O4 nanowires as high power cathode
7. Simon, P. & Gogotsi, Y. Materials for electrochemical capacitors. Nature Mater. materials for Li-ion batteries. Nano Lett. 10, 3852–3856 (2010).
7, 845–854 (2008). 37. Delmer, O., Balaya, P., Kienle, L. & Maier, J. Enhanced potential of
8. Zhai, Y. et al. Carbon materials for chemical capacitive energy storage. Adv. amorphous electrode materials: case study of RuO2 . Adv. Mater.
Mater. 23, 4828–4850 (2011). 20, 501–505 (2008).
9. Li, W., Dahn, J. R. & Wainwright, D. S. Rechargeable lithium batteries with 38. Zhong, K. et al. Investigation on porous MnO microsphere anode for lithium
aqueous electrolytes. Science 264, 1115–1118 (1994). ion batteries. J. Power Sources 196, 6802–6808 (2011).

6 NATURE ENERGY | VOL 1 | MAY 2016 | www.nature.com/natureenergy

© 2016 Macmillan Publishers Limited. All rights reserved


NATURE ENERGY DOI: 10.1038/NENERGY.2016.39 ARTICLES
39. Pan, H. et al. Sodium storage and transport properties in layered Na2 Ti3 O7 Author contributions
for room-temperature sodium-ion batteries. Adv. Energy Mater. Y.S. and J.L. proposed the research. H.P., Y.S. and J.L. designed the experiments. H.P. and
3, 1186–1194 (2013). Y.S. performed the material process, characterization, electrochemical measurements and
40. Divya, K. C. & Østergaard, J. Battery energy storage technology for power analysed the data. Y.C. synthesized the material. P.Y., Y.C. and C.W. conducted the TEM
systems—An overview. Electr. Power Syst. Res. 79, 511–520 (2009). and STEM mapping. K.S.H. and K.T.M. performed NMR characterization. H.P.,
Y.S. and J.L. co-wrote the paper. All authors discussed the results and commented on
the manuscript.
Acknowledgements
This work is supported by the US Department of Energy (DOE), Office of Basic Energy Additional information
Sciences, Division of Materials Sciences and Engineering, under Award Supplementary information is available online. Reprints and permissions information is
KC020105-FWP12152. The TEM, NMR and XRD work were performed using EMSL, a available online at www.nature.com/reprints. Correspondence and requests for materials
National Scientific User Facility sponsored by the Department of Energy’s Office of should be addressed to Y.S. or J.L.
Biological and Environmental Research and located at PNNL. PNNL is a Multi-Program
National Laboratory operated for DOE by Battelle. The work at UW was supported by Competing interests
Inamori Foundation. The authors declare no competing financial interests.

NATURE ENERGY | VOL 1 | MAY 2016 | www.nature.com/natureenergy 7


© 2016 Macmillan Publishers Limited. All rights reserved

You might also like