Você está na página 1de 9

View Online / Journal Homepage / Table of Contents for this issue

RSC Books
Downloaded on 14 February 2012
Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

Defining Chemical Science Content


www.rsc.org/materials

Volume 22 | Number 6 | 14 February 2012 | Pages 23172792

Cutting-edge high-quality content, outstanding excellence across the chemical sciences and beyond, the
RSC continues to lead as one of the fastest chemical sciences print and online book publishers in the world.
For postgraduates up to faculty members, our professional reference series provide authoritative coverage
in a comprehensive range of subjects including:

RSC Drug Discovery


ov
Series

RSC Biomolecular Sciences Series


R

RSC Gree
Green
reen
een C
Chemistry
Chemis Series

RSC Energy & Environment


ment Se
Series
eries

RSC Catalysis Series

RSC Nanoscience
os
&
Nanotec
Nanotechnology
not
lo Series

General
nerall Sc
Science

Seminal
mi Texts

Sign up
for
newsle book
tters.
Stay in
formed
.
GO
www.rs TO
c.org/a
lerts

For undergraduates and postgraduates, our textbooks will accompany


and support their studies. Popular science titles, ideal for the more
general reader, will enthuse and inform.

New Series - Coming Soon!


Polymer Chemistry, Smart Materials, NMR, New Developments Mass Spectrometry, Food & Nutritional Components in Focus

ISSN 0959-9428

www.rsc.org/books
PAPER
Hong Jin Fan et al.
Three-dimensional tubular arrays of MnO2NiO nanoflakes with high areal
pseudocapacitance

Registered Charity Number 207890

Journal of
Materials Chemistry

Dynamic Article Links <

Cite this: J. Mater. Chem., 2012, 22, 2419

PAPER

www.rsc.org/materials

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

Three-dimensional tubular arrays of MnO2NiO nanoflakes with high areal


pseudocapacitance
Jinping Liu,ab Jian Jiang,b Michel Bosmanc and Hong Jin Fan*ad
Received 27th September 2011, Accepted 7th November 2011
DOI: 10.1039/c1jm14804d
Transition metal oxide nanostructures are current research focus for energy storage applications. We
herein report the synthesis of MnO2NiO nanoflake-assembled tubular array on stainless steel
substrate to function as pseudocapacitor electrode by programmed three-dimensional (3D) interfacial
reactions, in which the ZnO nanowire array is employed as the low-cost in situ sacrificial template. In
this 3D nanoelectrode, MnO2 and NiO nanoflakes share the same root and form an integrated
hierarchical structure, which adheres robustly to the substrate. Importantly, both MnO2 and NiO
contribute to the charge storage. The highly porous structure, which allows easy penetration of the
electrolyte, gives additional merits. Detailed electrochemical characterization reveals that the
assembled MnO2NiO array exhibits good rate performance and cycle life. In particular, it displays an
areal capacitance that is four orders of magnitude higher than that of carbonaceous materials and
significantly superior to those of previous directly-grown pseudocapacitive nanostructure films.

1. Introduction
Driven by growing concerns about the depletion of traditional
energy resources and global warming caused by massive CO2
emissions, developing sustainable and clean energy products such
as electricity from solar cells and wind has attracted much interest
in recent years.15 Since most of the sustainable resources are only
intermittently available, to promote their full potential utilization,
it is highly required to explore advanced energy storage and
delivery systems.6,7 For electrochemical energy storage, a critical
element in the pursuit of this quest is the design of novel electrode
materials.810 Nanostructures have been demonstrated to be
unique in facilitating the mass transport, ion diffusion and electron transfer, thus dramatically boosting the electrochemical
performance.1113 In particular, ordered nanostructures grown
directly on current collectors are most helpful to charge carrier
transport and stress relaxation.1422 When combined with highly
porous or hollow structural morphologies,23 nanostructure arrays
represent an attractive architecture for fast reaction kinetics due
to significantly enhanced surface area. During the past decade,
chemical
synthesis
such
as
hydrothermal1822
and

a
Division of Physics and Applied Physics, School of Physical and
Mathematical Sciences, Nanyang Technological University, 21 Nanyang
Link, 637371, Singapore. E-mail: fanhj@ntu.edu.sg
b
Institute of Nanoscience and Nanotechnology, Department of Physics,
Central China Normal University, Wuhan, 430079, P. R. China
c
Institute of Materials Research and Engineering, Agency for Science,
Technology and Research (A*STAR), 3 Research Link, 117602,
Singapore
d
Energy Research Institute @ Nanyang Technological University
(ERI@N), 637553, Singapore

This journal is The Royal Society of Chemistry 2012

electrodeposition1416 demonstrated the impressive success in


achieving controllable composition and morphology of various
ordered electrode materials. However, for some purposes such as
optimizing electrode performance, it is desirable to combine
merits of individual materials.2426 In this respect, simple chemical
synthesis is not a fitting choice; it is still challenging to develop
a facile and general route for producing multi-component nanostructure arrays on large area electrode substrate.
Supercapacitors are one of the most promising candidates to
couple renewable energy sources for green transportation and
large-scale energy storage, owing to their various desirable
characteristics including long cycle life and operation at fast
charge and discharge rates.27 For example, conventional electric
double layer capacitors based on carbon typically have high
power density of >10 kW kg1 and very long lifetime of >106
cycles.7 However, due to the low density and gravimetric
capacitance of carbon, the capacitance per unit area (Ca) is
generally in the range of 1040 mF cm2.27 For applications such
as small scale electronics and stationary energy storage devices,
the amount of energy stored per unit area is a critical figure of
merit, more important than energy per unit mass. As a result,
metal oxides such as RuO2, MnO2, NiO, CoOx, and electronically conducting polymers, or their composites, have been
used2833 as they have increased gravimetric capacitance via
pseudocapacitive redox reactions and much larger density, thus,
it is possible to increase the Ca while keeping similar (or even
smaller) film thickness as that of conventional carbon-based
electrode. Nevertheless, a literature survey has indicated that the
Ca of so far reported directly-grown pseudocapacitive material
films/arrays is still in the mF cm2 region.27,3437 To potentially
increase the Ca, it is necessary to integrate two or more active
J. Mater. Chem., 2012, 22, 24192426 | 2419

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

materials with high redox electroactivity into one ordered


nanostructure. By doing this, more space in the electrode films
would be efficiently used to store energy and possible synergistic
effect between active materials can be utilized.24 With the help of
ordered array architecture, we can expect to increase the Ca
almost without sacrificing cycle life and rate performance.
MnO2 and NiO are the two most widely investigated pseudocapacitive materials due to their low cost, high redox activity,
environmentally benign nature and natural abundance.26,34 The
theoretical gravimetric capacitance of MnO2 and NiO is as high as
1370 and 2573 F g1, respectively. While the individual capacitive property of these two materials has been extensively demonstrated,22,2426,30,32,34,36,37 no study on their composites
electrochemical capacitance has been reported. In this paper, we
present a programmed three-dimensional (3D) interfacial reaction route to fabricate integrated MnO2NiO tubular arrays,
which are constructed uniformly by numerous thin nanoflakes.
The developed method can guarantee the successive deposition of
structural components and the in situ removal of the template in
the absence of any strong acid or alkali, with ease of structural
control and high generality. With the smart combination of MnO2
and NiO, a synergistic effect can be observed: the firstly-grown
MnO2 flake-assembled array provides a scaffold for the later NiO
growth, avoiding the conventional aggregation and ensuring
sufficient ion diffusion; while the further introduction of NiO
reduces the charge-transfer resistance of MnO2, leading to fast
electron transport within active materials. Consequently, the Ca
of 0.35 F cm2 can be achieved, four orders of magnitude higher
than the Ca values of carbonaceous materials (1040 mF cm2) and
significantly superior to those of most reported directly-grown
pseudocapacitive nanostructure films (0.5125 mF cm2). The
nanoflake-assembled tubular structures aligned directly on
current collector ensure the large electrochemically active area
and enhanced electron collection efficiency, also leading to good
rate capability and long cycle life.

2. Experimental
Synthesis
A typical fabrication procedure is shown in Scheme 1. We started
with ZnO nanowires, which are a common fabrication template
due to its easy availability and dissolution in acid or basic solution. Carbon-coated ZnO (ZnO@C) nanowire arrays on stainless
steel substrate (current collector) were prepared according to our
previous report.38 Typically, a thin layer carbon with 6 nm
thickness could be painted homogenously on the hydrothermally

grown nanowire surface. To fabricate our MnO2NiO hybrid


tubular arrays, a 3D interfacial reaction protocol was purposely
designed. At first, ZnO@C arrays were immersed into a 0.03 M
KMnO4 aqueous solution and sealed in a Teflon-lined stainless
steel autoclave at 160  C for 5 h (Step 1, interfacial reaction
between C and KMnO4 to obtain ZnOMnO2 core-shell array).
Then, the sample further reacted in a 20 mL aqueous solution
containing 0.5 g Ni(NO3)2 at 120  C for 12 h (interfacial reaction
of Ni3+ and ZnO). This process led to the coating of nickel
hydroxide onto the nanowire surface and the in situ dissolution
of ZnO (Step 2). An additional annealing process (350  C for 2 h)
in Ar gas was finally carried out to transform nickel hydroxide
into NiO as well as to improve the mechanical and electrical
adhesion between hybrid structure and current collector
substrate. For electrochemical property comparison, pure MnO2
tubular arrays without the second step of NiO growth were also
prepared by acid etching (0.1 M hydrochloric acid solution) of
ZnO@MnO2 core-shell arrays followed by 350  C annealing.
Structural characterization
The microstructure and morphology of the product were characterized using powder X-ray diffraction (XRD) (Bruker D-8
Avance), transmission electron microscopy (TEM) (JEM2010FEF, 200 kV), scanning electron microscopy (SEM) (JSM6700F, 5.0 kV), and Raman spectroscopy (Witech CRM200,
532 nm). Energy dispersive X-ray (EDX) analysis was conducted
on an FEI Titan (200 kV) TEM and acquired in scanning TEM
(STEM) mode at an equal acquisition time, with a nominal
electron beam diameter of 1 nm for the measurement. X-ray
photoelectron spectroscopy (XPS) measurement were performed
on a VG ESCALAB 250 spectrometer (Thermo Electron, U.K.)
with monochromatic Al Ka (1486.6 eV) irradiation.
Electrochemical test
Electrochemical measurements were performed on an electrochemical workstation (CHI 760C, CH Instruments Inc.,
Shanghai) using a three-electrode mode in 1.5 M LiOH aqueous
solution within the potential window of approximately 0.2 to
0.6 V. A piece of 2 cm2 MnO2NiO hybrid tubular array on
stainless steel foil was directly used as the working electrode. The
reference electrode and counter electrode were Ag/AgCl and
a platinum plate, respectively. EIS measurements were performed by applying an AC voltage with 5 mV amplitude in
a frequency range from 0.01 Hz to 100 kHz at open circuit
potential. Prior to the experiments, the array electrode was sealed

Scheme 1 Schematic illustration of the fabrication process of MnO2NiO tubular array electrode.

2420 | J. Mater. Chem., 2012, 22, 24192426

This journal is The Royal Society of Chemistry 2012

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

on all edges except for the working surface area with epoxy
resin. Areal capacitance was calculated using Equation: Carea
I$t/(OV$S), where I is the constant discharge current, t is the
discharging time, OV is the voltage drop upon discharging
(excluding the IR drop), and S is the geometrical area of the
electrode. The areal energy density de was calculated from the
galvanostatic chargedischarge curves according to Equation:
de Carea$(OV)2/2.

3. Results and discussion


Fig. 1a shows the XRD result of the ZnO@C array after the
reaction in KMnO4 solution. It is well known that carbon

(amorphous carbon, graphitic carbons such as carbon nanotube


and graphene) can easily have a redox reaction with KMnO4
even at room temperature, producing MnO2 materials with welldefined morphology (3C + 4MnO4 + H2O 4MnO2 + CO32 +
2HCO3).39,40 In our case, the carbon layers uniformly coated on
ZnO nanowires serve as a sacrificial template to direct the 3D
interfacial reaction. As shown in Fig. 1a, one characteristic peak
at 12.2 from layered birnessite-type MnO2 (JCPDS No. 801098) can be confirmed in addition to those from ZnO and
current collector substrate. Raman spectrum in Fig. 1b shows
that except for the peaks at 332 and 437 cm1 of single-crystalline
ZnO,38 the others are well indexed to MnO2. Raman bands at 655
and 577 cm1 can be recognized as the symmetric stretching

Fig. 1 (a) XRD result and (b) Raman spectrum of the array after the first 3D interfacial reaction. (c, d) Top-view SEM images of the ZnO@MnO2 coreshell nanoarray. Inset in Fig. 1d is the enlarged image of an individual structure. (e) Cross-sectional SEM image and (f) XPS results of the ZnO@MnO2
array.

This journal is The Royal Society of Chemistry 2012

J. Mater. Chem., 2012, 22, 24192426 | 2421

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

vibration (MnO) of the MnO6 groups and the (MnO)


stretching vibration in the basal plane of MnO6 sheet, respectively.40 No Raman signals of carbon were detected when the
recording wavelength was extended up to 2000 cm1, indicating
the complete consumption of carbon. The obtained MnO2@ZnO
core-shell nanowires are still well aligned on the substrate on
a large scale, as displayed in Fig. 1c. A close SEM examination
demonstrates the uniform coverage of MnO2 nanoflakes on ZnO
nanowire surfaces (Fig. 1d and inset). Cross-sectional image of
the array shown in Fig. 1e unambiguously reveals the entire and
intimate coating of each ZnO nanowire with flaky MnO2. To
investigate the oxidation state of Mn and further confirm the
composition, XPS of the array was recorded, as illustrated in
Fig. 1f. The Mn 2p, Mn 3s and O 1s spectra are individually
shown in the inset. The peak located at 529.8 eV is characteristic
of O element in oxide.41 The Mn 2p3/2 peak is centered at
642.4 eV and Mn 2p1/2 peak at 654.0 eV, with a spin-energy
separation of 11.6 eV, which are in good accordance with
previous data observed in MnO2.40,41 The Mn 3s spectrum shows
doublet peaks that result from the parallel spin coupling between
the electrons in 3s and 3d orbitals. The splitting width is 4.73 eV,
indicating that Mn oxidation state is 4 on the basis of an
approximately linear relationship between the Mn 3s splitting
widths and the Mn oxidation states.42
After the formation of MnO2@ZnO core-shell nanowire
array, the second 3D interfacial reaction was applied to fabricate
NiOMnO2 hybrid tubular array. This interfacial reaction is
enabled by the mutually promoted reactions of Ni2+ hydrolysis
and ZnO etching by H+ ions at the nanowire/solution interface. It
is well known that Ni2+ can hydrolyze and give rise to proton
(H+), especially under hydrothermal conditions. Since MnO2
nanoflakes grown on ZnO nanowires (Scheme 1, step 1) are
interpenetrated and form a highly porous structure, H+ is
believed to access nanowire/solution interface very easily, which
initiates the dissolution of ZnO nanowires. In turn, the
consumption of H+ by interfacial ZnO would accelerate the
hydrolysis of Ni2+, in this case, even the concentration of Ni2+ is
not as high as the general level for homogeneous nucleation,
nickel hydroxide could heterogeneously nucleate at the interface
and subsequently stretches out through the voids of MnO2
nanoflake networks. With the continuous dissolution of ZnO in
the vicinity of the interface, more and more nickel hydroxides
generate due to the accelerated Ni2+ hydrolysis, finally, nickel
hydroxideMnO2 tubular arrays can be obtained. After further
annealing, nickel hydroxide is readily converted into NiO, and
NiOMnO2 tubular arrays are attained accordingly. As displayed in Fig. 2a, NiO flexible nanoflakes densely cover on MnO2
and are interconnected crossing the ordered arrays, filling nearly
all the void spaces. The underlying MnO2 can still be identified
due to the contrast difference, as denoted by dashed circles in
Fig. 2b. The flaky morphology of NiO is inherited from nickel
hydroxide, which in general grows into two-dimensional nanostructures due to its layered crystal structure. Raman spectrum in
Fig. 2c reveals that in addition to the peak at 655 cm1 from
MnO2,40 the others can be well indexed to pure NiO. In detail, the
first three bands have vibrational origin and correspond to onephonon (1P) transverse optical (TO) and longitudinal optical
(LO) modes (at 515 cm1), two-phonon (2P) TO + LO
(at 1008 cm1) and 2LO (at 1153 cm1) modes. The last strong
2422 | J. Mater. Chem., 2012, 22, 24192426

Fig. 2 (a, b) SEM images of the array after the second 3D interfacial
reaction. Dashed circles in Fig. 2b reveal that the underlying tips of MnO2
can still be observed. (c) Raman spectrum of the MnO2NiO array.

band at 1514 cm1 is due to a two-magnon (2M) scattering.43 The


absence of ZnO Raman peaks reveals the successful dissolution
of ZnO single-crystal nanowires, as expected.
It should be pointed out that the second 3D interfacial reaction
occurs at the interface between the pre-grown MnO2 and ZnO
nanowires, thereby the post-grown NiO nanoflakes have the
same roots to MnO2 nanoflakes, which would make the hybrid
structure highly integrated. Considering that the carbon layertemplated growth of MnO2 is not limited to ZnO nanowire
arrays and a variety of hydrolyzable metal salts are available, the
combination of the above two interfacial reactions can potentially produce a wide spectrum of functional hybrid nanostructure arrays for various applications in catalysis, sensing and
energy conversion and storage, etc. One example could be MnO2/
a-Fe2O3 for lithium-ion battery anode material.
The structure and morphology evolution of the hybrid array
was further investigated by TEM. As illustrated in Fig. 3a, MnO2
layer of 50 nm thickness is deposited tightly on ZnO nanowire
after the carbon-templated interfacial reaction. It is noted that
some ZnO wires have already been etched at this growth stage,
possibly due to the presence of HCO3 in solution (produced
from the reaction between carbon and MnO4). Completely
tubular structure is observed after the second interfacial growth,
as shown in Fig. 3b. Even with two components integrated, the
structure is still highly uniform with straight and clear tube walls.
Enlarged TEM image (Fig. 3c and inset) from the tube surface
clearly demonstrates that NiO nanoflakes are porous (25 nm)
but structurally continuous. The pores are generated because of
the loss of water from the structure during the thermal decomposition of nickel hydroxide. The flake thickness is determined
less than 10 nm from the cross sections (see black arrows). Darkfield STEM image of two tubular structures is selected in Fig. 3d
for EDX analysis. Two different positions (P1: inner flakes; P2:
outer flakes) were detected and the results are shown in Fig. 3e.
Note that carbon and Cu signals are from the carbon-supported
Cu grid. While the inner nanoflakes contain both Ni and Mn
elements, the outer flakes are lack of Mn, in well coincidence with
our structure design protocol and further confirming that the
This journal is The Royal Society of Chemistry 2012

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

Fig. 3 (a) TEM image of ZnOMnO2 core-shell nanowires. One tube is


also seen here. (b) TEM image of one MnO2NiO branched tube.
(c) Enlarged images of several NiO nanoflakes from the hybrid tube. The
arrows indicate the cross section of NiO flakes. (d) Dark-field STEM
image of two tubular structures. (e) EDX analysis of two positions
(P1, P2) shown in Fig. 3d.

growth of nickel hydroxide started from the MnO2/ZnO interface. The presence of Zn traces should be ascribed to the residual
Zn species, but it does not affect the electrochemical performance
of hybrid structure.
To test the electrochemical capacitive performance of the
MnO2NiO tubular array, cyclic voltammogram (CV) was firstly
recorded at the scan rate of 50 mV s1 (Fig. 4a). The CV of
pristine MnO2 tubular array is also shown for comparison. The

CV measurements were conducted in a three-electrode system


using saturated Ag/AgCl as the reference and platinum foil as the
counter-electrode. For pure MnO2, the CV curve exhibits rectangular-like shape without obvious redox peaks; the capacitive
behavior of MnO2 can be ascribed to surface cation (Li+)
adsorption from electrolyte and possible intercalation/deintercalation of Li+.26 With the NiO integrated, the CV of hybrid
array demonstrates distinct difference. While there is no obvious
CV shape change within the potential window of 0.20.2 V,
a pair of well-defined redox peaks at 0.33 and 0.46 V can
be clearly observed, which are due to the reversible reactions of
Ni2+/Ni3+ associated with anions OH: NiO + OH 4 NiOOH +
e.30,34 The CV integrated area of hybrid array is apparently
larger than that of pristine MnO2 array; this will lead to significant increase in the pseudocapacitance. The specific capacitances
of both MnO2NiO and MnO2 tubular arrays were further
calculated from charge-discharge profiles. Fig. 4b presents their
chargedischarge curves in the potential range of 0.20.6 V at
current density of 8.5 mA cm2. The voltage plateaus for MnO2
NiO array match well with the peaks observed in the CV curve.
The charging/discharging times are nearly the same, indicating
superior reversibility of the redox reaction. The Carea is further
estimated to be 0.35 F cm2, which is nearly 3.5 times that of our
pristine MnO2 tubular array (0.101 F cm2). This value is also
4 orders of magnitude higher than those of conventional
carbonaceous materials (1040 mF cm2) and much greater than
previously reported results of most directly-grown pseudocapacitive nanostructure films such as NiOTiO2 nanotube array
(2.54 mF cm2),34 NiNiO core shell inverse opal film
(10 mF cm2),37 TiO2 nanotube film (0.5380.911 mF cm2)27
and electrodeposited MnO2 nanosheet film (0.125 F cm2).36 The
high Carea of our hybrid array results from the effective
integration of nanoflake-shaped MnO2 and NiO into an ordered

Fig. 4 Electrochemical performance comparison of MnO2 and MnO2NiO tubular arrays. (a) CVs at 50mV s1. (b) Chargedischarge curves of the first
cycle at the current density of 8.5 mA cm2. (c) EIS results. (d) Cycling performance at current rate of 8.5 mA cm2.

This journal is The Royal Society of Chemistry 2012

J. Mater. Chem., 2012, 22, 24192426 | 2423

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

and tubular structure directly on current collector. In detail, the


constituent high-capacitance pseudocapacitive materials grow
densely on entire electrode surface without electrochemicallyinert additives; this not only saves the active spaces but might
also offer greatly improved electron transfer pathways.
Furthermore, the porous nature of NiO nanoflakes and interspacing void formed by interconnected flakes as well as the
tubular structure can allow an effective contact between the
electrolyte (Li+, OH) and the two active materials, which
guarantees adequate ion supply for redox reactions.
To further understand the fundamental behavior of supercapacitor electrodes, EIS analysis was measured and the corresponding Nyquist plots are shown in Fig. 4c. The impedance
spectra of hybrid MnO2NiO and pure MnO2 arrays are almost
similar in form with a semicircle at a higher frequency region and
a spike at lower frequency. EIS data were analyzed using the
complex nonlinear least-squares (CNLS) fitting method26a on the
basis of the equivalent circuit consisting of a combined resistance
Re (ionic resistance of electrolyte, intrinsic resistance of substrate,
and contact resistance at the active material/current collector
interface), a charge-transfer Rct, Warburg resistance Zw, and
double-layer capacitance Cdl on the grain surface (inset of Fig. 4c).
Re and Rct can be obtained from the Nyquist plots, where the high
frequency semicircle intercepts the real axis at Re and (Re + Rct),
respectively. The slope of the 45 portion of the curve at low
frequency is Zw, which is a result of the frequency dependence of
ion diffusion/transport in the electrolyte to the electrode surface.24
Fig. 4c shows that Re and Zw values are almost the same for both
arrays; a major difference is the charge-transfer resistance Rct. The
Rct values are calculated to be 18.7 and 54 U for MnO2NiO and
pristine MnO2 tubular arrays, respectively. This clearly demonstrates that more facile charge transfer is achieved by integrating
NiO nanoflakes into the tubular array architecture. The reduced
Faraday resistance rendered by the intriguing material combination and the hierarchical structure leads to enhanced electrochemical reaction, contributing greatly to the observed large Carea
value of MnO2NiO hybrid array.
In addition to the much higher Ca, the MnO2NiO hybrid
array also demonstrates superior cycling performance to pristine
MnO2 array. Fig. 4d illustrates the discharge capacitance of the
above two arrays as a function of cycle number at the current
density of 8.5 mA cm2. The hybrid array exhibits excellent
electrochemical stability with only 3.6% deterioration of the
initial capacitance after 1500 cycles. The pristine MnO2 array,

however, has 9.2% capacitance loss after cycling. It is believed


that the intimate root contact of NiO flakes with MnO2 flakes
and the networking of MnO2 by filling in the array voids with
flexible interpenetrated NiO flakes can help to maintain the
structural integrity and mechanical adhesion with current
collector, thus be beneficial to long-term electrochemical cycling.
Good rate capability is an important requirement for supercapacitors. The MnO2NiO hybrid array was purposely tested at
different current densities and the galvanostatic discharge curves
are shown in Fig. 5a. The Carea value versus current density plots
are also presented in Fig. 5b, in which the times required for full
discharging at each current density are indicated. Since the
reversible redox reaction is a highly diffusion-controlled process,
the capacitance will decrease at fast charge-discharge rate. The
Carea at current densities of 5, 8.5, 15, 20, and 25 mA cm2 is
determined 0.4, 0.35, 0.3, 0.25, 0.22 F cm2, respectively. At the
current density as high as 25 mA cm2 (discharge within 8.8 s),
55% of the capacitance at 5 mA cm2 can be retained, still much
higher than state-of-the-art carbonaceous materials and most
pseudocapacitive nanostructure arrays.27,34,36,37 We tested the
electrochemical performance of pure MnO2 tubular array at
higher current densities, but found that its capacitance decreased
dramatically when the current larger than 10 mA cm2 (data not
shown). The superior rate capability of MnO2NiO hybrid
tubular array can be attributed to its substantially reduced
charge-transfer resistance (ensuring fast redox kinetics) as evidenced by EIS. In Fig. 5b the areal energy density (de) of hybrid
array is also plotted versus the current density. The hybrid
tubular array consistently demonstrates energy density with the
order of magnitude of 105 Wh cm2. The de at 5 and 25 mA cm2
is 3.56 and 1.96  105 Wh cm2, respectively, which are several
times that of some carbon fiber and carbon nanotube-based
supercapacitors.44,45
To investigate the electrochemical performance of hybrid
array at high rates, we further tested the cycling performance at
a much larger current density of 15 mA cm2 (in general, 510
times that of commonly used value); the result is shown in
Fig. 6a. After 1500 cycles, the tubular hybrid array still retains
87.5% of the initial capacitance with >96% Coulombic efficiency
from the 10th cycle (finally 99%). The last 20 chargedischarge
curves are also displayed in Fig. 6b, which exhibit almost the
same shape, revealing excellent long-term cyclability of the
hybrid array electrode even under fast chargedischarge
conditions.

Fig. 5 (a) Discharge curves of MnO2NiO tubular array at different current densities. (b) Rate capability and energy density versus current density plot
for MnO2NiO array.

2424 | J. Mater. Chem., 2012, 22, 24192426

This journal is The Royal Society of Chemistry 2012

Notes and references


1
2
3
4
5

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Fig. 6 (a) Cycling stability of MnO2NiO tubular arrays at higher
current density of 15 mA cm2. Coulombic efficiency as the function of
cycle number is also plotted. (b) The chargedischarge curves of the last
20 cycles for the hybrid array.

21
22
23

4. Conclusions
In summary, MnO2NiO tubular arrays assembled by thin
nanoflakes on current collector have been successfully fabricated
based on two-step 3D interfacial reactions using ZnO nanowire
arrays as the sacrificial template. The proposed interfacial reaction route is general and might be easily extendable to synthesize
various functional hybrid nanostructure arrays. As a demonstration of the application potential, the electrochemical
performance of such a hybrid array as supercapacitor electrode
was systematically investigated. The results show that MnO2
NiO tubular arrays exhibit much higher areal capacitance than
conventional carbonaceous materials and many directly-grown
pseudocapacitive nanostructure films; they also displayed good
cyclability and rate capability. Our present work indicates that
integrated tubular arrays constructed by two or more highcapacitance pseudocapacitive materials hold great promise in
thin-film supercapacitors and other kinds of electrochemical
devices.

24
25
26
27
28
29
30
31
32
33
34
35
36
37

Acknowledgements

38

This work is partially supported by NAP start-up fund to H.J.F


by Nanyang Technological University (M5800014), Singapore
and the National Natural Science Foundation of China (No.
51102105) to J.P.L.

39

This journal is The Royal Society of Chemistry 2012

40
41

J.-M. Tarascon and M. Armand, Nature, 2001, 414, 359.


P. Simon and Y. Gogotsi, Nat. Mater., 2008, 7, 845.
C. Liu, F. Li, L. P. Ma and H. M. Cheng, Adv. Mater., 2010, 22, E28.
L. Li, H. Q. Wang, X. S. Fang, T. Y. Zhai, Y. Bando and D. Golberg,
Energy Environ. Sci., 2011, 4, 2586.
D. V. Esposito, S. T. Hunt, A. L. Stottlemyer, K. D. Dobson,
B. E. McCandless, R. W. Birkmire and J. G. Chen, Angew. Chem.,
Int. Ed., 2010, 49, 9859.
P. G. Bruce, B. Scrosati and J.-M. Tarascon, Angew. Chem., Int. Ed.,
2008, 47, 2930.
J. R. Miller and P. Simon, Science, 2008, 321, 651.
M. G. Kim and J. Cho, Adv. Funct. Mater., 2009, 19, 1497.
Y. G. Guo, J. S. Hu and L. J. Wan, Adv. Mater., 2008, 20,
2878.
J. Liu, G. Z. Cao, Z. G. Yang, D. H. Wang, D. Dubois, X. D. Zhou,
G. L. Graff, L. R. Pederson and J. G. Zhang, ChemSusChem, 2008, 1,
676.
M. D. Stoller and R. S. Ruoff, Energy Environ. Sci., 2010, 3, 1294.
G. Lota, K. Fic and E. Frackowiak, Energy Environ. Sci., 2011, 4,
1592.
E. Kang, Y. S. Jung, A. S. Cavanagh, G.-H. Kim, S. M. George,
A. C. Dillon, J. K. Kim and J. Lee, Adv. Funct. Mater., 2011, 21, 2430.
P. L. Taberna, S. Mitra, P. Poizot, P. Simon and J. M. Tarascon, Nat.
Mater., 2006, 5, 567.
C. R. Sides and C. R. Martin, Adv. Mater., 2005, 17, 125.
G. Che, K. B. Jirage, E. R. Fisher and C. R. Martin, J. Electrochem.
Soc., 1997, 144, 4296.
C. K. Chan, H. Peng, G. Liu, K. McIlwrath, X. F. Zhang,
R. A. Huggins and Y. Cui, Nat. Nanotechnol., 2008, 3, 31.
Y. Li, B. Tan and Y. Wu, Nano Lett., 2008, 8, 265.
J. P. Liu, C. W. Cheng, W. W. Zhou, H. X. Li and H. J. Fan, Chem.
Commun., 2011, 47, 3436.
(a) J. P. Liu, Y. Y. Li, X. T. Huang, G. Y. Li and Z. K. Li, Adv. Funct.
Mater., 2008, 18, 1448; (b) G. W. Ho and A. S. W. Wong, Appl. Phys.
A: Mater. Sci. Process., 2007, 86, 457.
J. P. Liu, Y. Y. Li, X. T. Huang, R. M. Ding, Y. Y. Hu, J. Jiang and
L. Liao, J. Mater. Chem., 2009, 19, 1859.
X. H. Lu, D. Z. Zheng, T. Zhai, Z. Q. Liu, Y. Y. Huang, S. L. Xie and
Y. X. Tong, Energy Environ. Sci., 2011, 4, 2915.
J. P. Liu, Y. Y. Li, H. J. Fan, Z. H. Zhu, J. Jiang, R. M. Ding,
Y. Y. Hu and X. T. Huang, Chem. Mater., 2010, 22, 212.
J. P. Liu, J. Jiang, C. W. Cheng, H. X. Li, J. X. Zhang, H. Gong and
H. J. Fan, Adv. Mater., 2011, 23, 2076.
Y. B. He, G. R. Li, Z. L. Wang, C. Y. Su and Y. X. Tong, Energy
Environ. Sci., 2011, 4, 1288.
(a) L. H. Bao, J. F. Zang and X. D. Li, Nano Lett., 2011, 11, 1215; (b)
H. Jiang, L. P. Yang, C. Z. Li, C. Y. Yan, P. S. Lee and J. Ma, Energy
Environ. Sci., 2011, 4, 1813.
M. Salari, S. H. Aboutalebi, K. Konstantinov and H. K. Liu, Phys.
Chem. Chem. Phys., 2011, 13, 5038.
C. C. Hu, K. H. Chang, M. C. Lin and Y. T. Wu, Nano Lett., 2006, 6,
2690.
G. Yang, C. Xu and H. Li, Chem. Commun., 2008, 6537.
T. Zhu, J. S. Chen and X. W. Lou, J. Mater. Chem., 2010, 20, 7015.
H. Zhang, G. P. Cao, Z. Y. Wang, Y. S. Yang, Z. J. Shi and Z. N. Gu,
Nano Lett., 2008, 8, 2664.
R. Liu and S. B. Lee, J. Am. Chem. Soc., 2008, 130, 2942.
L. Gao, F. Xu, Y. Y. Liang and H. L. Li, Adv. Mater., 2004, 16, 1853.
J.-H. Kim, K. Zhu, Y. F. Yan, C. L. Perkins and A. Frank, Nano
Lett., 2010, 10, 4099.
N. K. Shrestha, Y.-C. Nah, H. Tsuchiya and P. Schmuki, Chem.
Commun., 2009, 2008.
W. Xiao, H. Xia, J. Y. H. Fuh and L. Lu, J. Electrochem. Soc., 2009,
156, A627.
J. H. Kim, S. H. Kang, K. Zhu, J. Y. Kim, N. R. Neale and
A. J. Frank, Chem. Commun., 2011, 47, 5214.
J. P. Liu, Y. Y. Li, R. M. Ding, J. Jiang, Y. Y. Hu, Q. B. Chi,
Z. H. Zhu and X. T. Huang, J. Phys. Chem. C, 2009, 113, 5336.
S. W. Lee, J. Kim, S. Chen, P. T. Hammond and S. H. Yang, ACS
Nano, 2010, 4, 3889.
J. Yan, Z. J. Fan, T. Wei, W. Z. Qian, M. L. Zhang and F. Wei,
Carbon, 2010, 48, 3825.
H. Xia, M. O. Lai and L. Lu, J. Mater. Chem., 2010, 20, 6896.

J. Mater. Chem., 2012, 22, 24192426 | 2425

44 C. Du, J. Yeh and N. Pan, Nanotechnology, 2005, 16, 350.


45 C. Niu, E. K. Sichel, R. Hoch, D. Moy and H. Tennent, Appl. Phys.
Lett., 1997, 70, 1480.

Downloaded on 14 February 2012


Published on 13 December 2011 on http://pubs.rsc.org | doi:10.1039/C1JM14804D

42 M. Chigane and M. Ishikawa, J. Electrochem. Soc., 2000, 147, 2246.


43 N. M. -Ulmane, A. Kuzmin, I. Steins, J. Grabis, I. Sildos and M. Pars,
J. Phys. Conf. Ser., 2007, 93, 012039.

2426 | J. Mater. Chem., 2012, 22, 24192426

This journal is The Royal Society of Chemistry 2012

Você também pode gostar