ArticlePDF Available

Nanoelectrode design from microminiaturized honeycomb monolith with ultrathin and stiff nanoscaffold for high-energy micro-supercapacitors

Springer Nature
Nature Communications
Authors:

Abstract and Figures

Downsizing the cell size of honeycomb monoliths to nanoscale would offer high freedom of nanostructure design beyond their capability for broad applications in different fields. However, the microminiaturization of honeycomb monoliths remains a challenge. Here, we report the fabrication of microminiaturized honeycomb monoliths—honeycomb alumina nanoscaffold—and thus as a robust nanostructuring platform to assemble active materials for micro-supercapacitors. The representative honeycomb alumina nanoscaffold with hexagonal cell arrangement and 400 nm inter-cell spacing has an ultrathin but stiff nanoscaffold with only 16 ± 2 nm cell-wall-thickness, resulting in a cell density of 4.65 × 109 cells per square inch, a surface area enhancement factor of 240, and a relative density of 0.0784. These features allow nanoelectrodes based on honeycomb alumina nanoscaffold synergizing both effective ion migration and ample electroactive surface area within limited footprint. A micro-supercapacitor is finally constructed and exhibits record high performance, suggesting the feasibility of the current design for energy storage devices. Micro-supercapacitors are promising energy storage systems to power the future electronic devices. Here, the authors utilize honeycomb alumina nanoscaffold as a nanostructuring platform to design nanoelectrodes and construct micro-supercapacitors with impressive performance.
This content is subject to copyright. Terms and conditions apply.
ARTICLE
Nanoelectrode design from microminiaturized
honeycomb monolith with ultrathin and stiff
nanoscaffold for high-energy micro-
supercapacitors
Zhendong Lei1,2,5, Long Liu1,5, Huaping Zhao1*, Feng Liang 3*, Shilei Chang3, Lei Li4, Yong Zhang 2,
Zhan Lin4*, Jörg Kröger 1& Yong Lei 1*
Downsizing the cell size of honeycomb monoliths to nanoscale would offer high freedom of
nanostructure design beyond their capability for broad applications in different elds. How-
ever, the microminiaturization of honeycomb monoliths remains a challenge. Here, we report
the fabrication of microminiaturized honeycomb monolithshoneycomb alumina nanoscaf-
foldand thus as a robust nanostructuring platform to assemble active materials for micro-
supercapacitors. The representative honeycomb alumina nanoscaffold with hexagonal cell
arrangement and 400 nm inter-cell spacing has an ultrathin but stiff nanoscaffold with only
16 ± 2 nm cell-wall-thickness, resulting in a cell density of 4.65 × 109cells per square inch, a
surface area enhancement factor of 240, and a relative density of 0.0784. These features
allow nanoelectrodes based on honeycomb alumina nanoscaffold synergizing both effective
ion migration and ample electroactive surface area within limited footprint. A micro-
supercapacitor is nally constructed and exhibits record high performance, suggesting the
feasibility of the current design for energy storage devices.
https://doi.org/10.1038/s41467-019-14170-6 OPEN
1Institute of Physics and IMN MacroNano®, Ilmenau University of Technology, Ilmenau, Germany. 2NUS Graduate School for Integrative Sciences and
Engineering, National University of Singapore, Singapore, Singapore. 3Faculty of Metallurgical and Energy Engineering, Kunming University of Science and
Technology, Kunming, China. 4School of Chemical Engineering and Light Industry, Guangdong University of Technology, Guangzhou, China.
5
These authors
contributed equally: Zhendong Lei, Long Liu. *email: huaping.zhao@tu-ilmenau.de;liangfeng@kust.edu.cn;zhanlin@gdut.edu.cn;yong.lei@tu-ilmenau.de
NATURE COMMUNICATIONS | (2020) 11:299 |https://doi .org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications 1
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Honeycomb monoliths (HMs) are a kind of cellular mate-
rials with two-dimensional cell conguration consisting of
parallel and straight channels extended throughout the
body1. HMs have been widely exploited as catalyst supporters in
different gaseous reactor applications such as chemical and
rening processes, catalytic combustion, ozone abatement, and
photocatalytic air purication27. As catalyst supporters, HMs
have higher surface-to-volume ratio comparing to that of planar
substrates for providing high contact efciencies between the
supported catalysts and the gaseous reactants, meanwhile their
parallel and straight channels facilitate mass transport of gas with
low diffusion resistance. Besides these gas-phase catalysis appli-
cations, the cellular and homogeneous structure of the HM pre-
sents potential opportunities for other functional applications.
For example, it could be a desirable platform to assemble an
electroactive electrode for electrochemical devices (e.g., batteries,
supercapacitors, fuel cells, water electrolyzer, and electrochemical
sensors), in which an efcient electrochemical reaction requires
both high specic surface area of the electrode and favorable ionic
transport within the electrode811. Despite these merits, so far
HMs have not been widely utilized for electrochemical applica-
tions mainly due to the bulky structure of the existing HMs. The
present electrode design based on HMs makes electrochemical
devices cumbersome (especially the device volume), and mean-
while the macrostructural channels of HMs require an imprac-
tically large amount of electrolytes to ll up all the channels in
order to ensure a sufcient contact between the supported elec-
troactive materials and the electrolyte, which is different from
gas-phase catalysis applications. Moreover, the specic surface
area of the conventional HMs, although it is higher than that of
planar substrate counterparts, is far from sufcient to satisfy the
requirements of electrochemical applications. Currently, the best
reported HM can only attain 1.6 × 104cells per square inch (cpsi)
and about 3 μm in cell wall thickness12, with a specic surface
area not being comparable with that of widely used nanoelec-
trodes (i.e., mainly arrays of nanowires or nanotubes). However,
most of those nanoelectrodes suffer from the limit of aspect ratio
to maintain their highly oriented nature by avoiding the
agglomeration since nanowires and nanotubes with high aspect
ratio would prefer to form into many dense clusters, and con-
sequently are difcult to simultaneously satisfy both high specic
surface area and low ion transport resistance. To this end,
downsizing the cell size (e.g., channel diameter and cell wall
thickness) of the conventional HMs to nanoscale shall be a
solution to efciently address the challenges for extending the
application potentials of HMs. Nevertheless, the micro-
miniaturization of the HMs encounters technological difculties
in creating exactly parallel and straight nanoscale channels over a
large area by all reported HM fabrication approaches whatever in
industry or in laboratory, and meanwhile faces the challenge to
guarantee that the microminiaturized HM would hold the similar
excellent mechanical stabilities as the conventional HM.
In this article, microminiaturized HMs with ultrathin and stiff
nanoscaffoldhoneycomb alumina nanoscaffold (HAN)is
realized for the rst time. Specically, a large-scale HAN is fab-
ricated by using a nanoindentation-anodization-etching process
with a high-purity aluminum foil. A representative HAN is
actually a microminiaturized HM with hexagonal cell arrange-
ment, 400 nm inter-cell spacing and only 16 ± 2 nm cell wall
thickness. Impressively, the cell density of the HAN reaches
4.65 × 109cpsi that is ve orders of magnitude higher than that of
the reported highest value. Meanwhile, the HAN with such
ultrathin cell wall is stiff and can sustain surprisingly high
mechanical stability identied by an experimental nanoindenta-
tion. Such robust HAN offers a stable nanostructuring platform
to assemble electroactive materials for micro-supercapacitors
(MSCs). The insulating HAN is retained in the nanoelectrodes
and thus endows the nanoelectrodes with vertically aligned and
robustly stable nanoporous structure to simultaneously achieve
both effective ion migration and ample electroactive surface area
within limited footprint. As a result, a MSC constructed with the
HAN-based nanoelectrodes exhibits record high-energy perfor-
mance among the reported micro-supercapacitors. The max-
imum capacitance of the MSC reaches 128 mF cm2at a current
density of 0.5 mA cm2, and the peak energy and power densities
are 160 μWh cm2and 40 mW cm2, respectively.
Results
Formation and structure of HAN. A schematic of the farbication
process of HAN, and the corresponding scanning electron
microscope (SEM) images, are shown in Fig. 1ah. Briey, the
fabrication of HAN includes nanoindentation, anodization, and a
two-step etching process (Fig. 1a). The fabrication process begins
with the anodization of a surface-nanopatterned aluminum sub-
strate to obtain nanoporous alumina (Al
2
O
3
) with a hexagonal
cell arrangement and a cell periodicity of 400 nm. Notablely, the
as-prepared nanoporous alumina has a double-layer cell structure
(Fig. 1b), with an acid acid anion-contaminated Al
2
O
3
thick layer
(in dark-gray color with the thickness of 90 ± 2 nm) adjacent to
the cell center and a relatively pure Al
2
O
3
thin layer (in light-gray
color with the thickness of 21 ± 2 nm) remote from the cell
center12. By applying a precisely controlled two-step etching in
aqueous H
3
PO
4
solutions, the acid anion-contaminated Al
2
O
3
thick layer are easy to be totally dissolved and meanwhile the pure
Al
2
O
3
thin layer is partially etched. Finally, a honeycomb
nanoscaffold consisting of only an ultrathin layer of pure Al
2
O
3
is
left with the cell wall thickness of 16 ± 2 nm (Fig. 1c), which is a
HAN (i.e., a microminiaturized HM). Figure 1d schematically
shows the evolution of the HAN through the second etching
process. In the case of the HAN with hexagonal cell arrangement
and 400 nm inter-cell spacing, the cell density is calculated to be
4.65 × 109cpsi that is ve orders of magnitude higher than that of
the reported highest value13. The cross-sectional SEM image of a
HAN with cell depth of about 25 μm conrms the straight and
stable nanoporous structure of the HAN (Fig. 1e). Noting that the
cell depth of the HAN depends on the anodization time, by
tuning the anodization time, the cell depth of the HAN can be
adjusted from hundreds of nanometers to hundreds of micro-
meters. The surface area enhancement factor for an above-
mentioned HAN with the 25-μm-deep cell is calculated to be 240.
And, importantly, the relative density of HAN is calculated to be
only 0.0784, corresponding to its porosity of 0.9216. When
applying such HAN as a scaffold for designing electrodes and/or
devices, the high specic surface area and the lower dead
volume(i.e., the proportion of the insulating Al
2
O
3
to the whole
electrode) in electrodes/devices will be simultaneously satised,
beneting to accomplishing high device performance. Not limited
to hexagonal cell arrangement with 400 nm inter-cell spacing, two
other HANs also have been produced through tuning the nano-
patterns on the surface of aluminum foils generated by nanoin-
dentation, one with hexagonal cell arrangement but 800 nm inter-
cell spacing (Fig. 1f) and the other with 400 nm inter-cell spacing
but square cell arrangement (Fig. 1g). Note that the HAN can be
fabricated over large area with almost no structural defects
(Fig. 1h and Supplementary Fig. 1ab), which is also an impor-
tant advantage of the HAN for electrode designing.
Generally, a compressive stack pressure typically in the range of
0.110 MPa is applied during the assembly of batteries (coin- and
punch-cell) and supercapacitors for the purpose of maintaining
intimate contact between the electrodes and the current collectors as
well as preventing active materials delamination and deformation in
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6
2NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
the electrodes during operation14. Obviously, such high extrusion
pressure would bring forward the critical challenges and require-
ments to the mechanical stabilities of the nanoelectrodes for
supercapacitors and batteries. As known, the cellular solids with
honeycombs are much stiffer and stronger to manifest high
compressive strength when loaded at the out-of-plane direction (i.e.,
along the cell axis) than at the in-plane direction15,16,thus
nanoindentation experiments have been further conducted for
characterizing the nanomechanical properties of the HAN. The
Youngs modulus of the HAN with hexagonal cell arrangement,
400 nm inter-cell spacing and only 16 nm cell wall thickness are
obtained in the range of 1.413.22 GPa measured at ve different
positions (Supplementary Fig. 1c). The excellent mechanical
performance of the HAN should be attributed to not only the
high hardness of the pure Al
2
O
3
and also the anisotropic
honeycomb microarchitectures of the HAN1720, and it would be
asignicant advantage when applying as nanostructuring platform
to assemble electroactive materials especially for supercapacitors
and batteries, because the HAN can play the part of a rigid keel in
the nanoelectrodes to sustain the mechanical extrusion during the
device assembly but maintain the structural features at the
nanoscale of the electrodes, which are inherited from the HAN,
for nally accomplishing nanoelectrodes with both high specic
surface area and low ion transport resistance for efcient
electrochemical energy storage.
Fabrication and characterizations of HAN@SnO
2
as nanos-
tructured current collectors. As aforementioned, the HAN is a
kind of microminiaturized HMs with much higher cell density
than the conventional HMs, and the cellular and homogeneous
structure of the HAN presents opportunities as nanostructuring
platform to assemble electroactive materials for electrochemical
devices. Given the insulating nature of the HAN, the rst step to
design nanoelectrodes with the HAN for electrochemical appli-
cations is to convert the insulating HAN to be a nanostructured
current collector. As shown in Fig. 2a, a 12-nm-thick tin oxide
(SnO
2
) layer was conformally coated onto the HAN with hex-
agonal cell arrangement and 400 nm inter-cell spacing (Fig. 1c)
through atomic layer deposition (ALD). Note that a thick layer of
nickel (~10 μm) as the supporting substrate was electrochemically
deposited onto the top of the HAN before the conduction of the
ALD process. The SnO
2
-coated HAN (denoted as HAN@SnO
2
)
preserves the original honeycomb features with the cell wall
thickness increasing from 16 ± 2 to 40 ± 2 nm. To verify the
possibility of HAN@SnO
2
as nanostructured current collectors, a
Nanoindentation
a
d
b
e
h
c
fg
Anodization 1st etching 2nd etching
2nd etching
Relatively
pure Al2O3
Acid anion
contaminated
Al2O3
Fig. 1 Fabrication and structure of HAN. a Illustration of the HAN fabrication process. bSEM image of nanoporous alumina after anodization process
revealing the double-layer cell structure. cSEM image of HAN with hexagonal cell arrangement and 400 nm inter-cell spacing showing the structure of
HAN with ultrathin wall of 16 ± 2 nm. dSchematic of the evolution of the HAN by the second etching process. eCross-sectional view SEM image of an
HAN indicating the cell depth to be about 25 μm. fSEM image of HAN with hexagonal cell arrangement and 800 nm inter-cell spacing. The wall thickness
of the cell is 18 ± 2 nm. gSEM image of HAN with square cell arrangement and 400 nm inter-cell spacing. The wall thickness of the cell is 12 ± 2 nm. h
Large-scale SEM image showing the uniform, stable and defect-free structure of the HAN with hexagonal cell arrangement and 400 nm inter-cell spacing.
Scale bar: 500 nm (b); 500 nm (c); 5 μm(e); 1 μm(f); 500 nm (g); 10 μm(h).
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6 ARTICLE
NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
symmetric supercapacitor based on two HAN@SnO
2
electrodes,
HAN@SnO
2
//HAN@SnO
2
, were assembled and characterized
since SnO
2
is also one kind of electroactive materials for super-
capacitors. Rate performance of supercapacitors, especially under
high scan rate, could indirectly reect the ionic and electronic
transport behavior of electrodes. The cyclic voltammetry (CV)
curves of the device at different scan rates are shown in Fig. 2b,
and all have a quasi-rectangular shape even at a high scan rate of
20 V s1, which indicates the typical double-layer capacitive
behavior of the device. Figure 2c reveals the corresponding var-
iation of the discharge current density at different scan rates.
Clearly, the discharge currents keep a linear relationship upon
scan rates nearly until 20 V s1, suggesting the high instantaneous
power characteristics of the device. The representative Nyquist
plot of the device, and the corresponding expanded view at the
high-frequency region, again reveals the capacitive behavior even
at high frequencies that are attributed to the fully accessible
surface area of the electrodes for electrolyte ions adsorption/
desorption (Supplementary Fig. 2). Moreover, the high-frequency
semicircle reects the small charge transfer resistance (R
ct
) of only
1.8 Ωin HAN@SnO
2
//HAN@SnO
2
device, also contributing to
realize high-rate performance. Therefore, high-rate performance
should be attributed not only to the rapid mass transport of ions
but also to the efcient charge transport and collection in
HAN@SnO
2
nanoelectrodes during the ultrafast charge-
discharge process21.
In addition, a nearly 98% of the initial performance was
retained after continued 30,000 charge/discharge cycles at a scan
rate of 1 V s1(Fig. 2d) and no obvious structural changes and/or
damages can be found for the electrodes when comparing the
SEM images of the HAN@SnO
2
electrodes before and after long-
termed cycling (Fig. 2a, e, respectively). The high stability in both
electrochemical performance and electrode structure should be
attributed to the existence of the HAN in the nanoelectrodes to
functionalize as mechanically robust keel. On the other hand, it is
well-known that the mechanical extrusion is an essential step
during the device assembly of supercapacitors and batteries14,
however, at the same time it would generate a destructive impact
on nanoelectrodes. For example, the collapse of nanostructures
and then the block of ionic transport pathway will subsequently
result in unexpectedly device performance degradation compared
to the result of three-electrode conguration. Attributing to the
high stiffness of Al
2
O
3
, the monolithic structure of HAN without
structural defects over a large area and the high Youngs modulus
a
d
g
e
h
f
i
b60
100
80 2
1
0
–1
–2
0.0 0.2 0.4 0.6 0.8 1.0
60
40
20
0
12
8
4
0
0
0
0.0 0.2 0.4 0.6
Potential (V)
Potential (V)
@ 500 mV s–1
@1 V s–1
0.8 1.0
4 8 12 16
1.5
0.5
–0.5
5000 10,000 15,000
Cycle number
Time (min)
20,000 25,000 30,000
40
30
20
10
0
0 5 10 15
6 MPa
10 MPa
8 MPa
Glass fiber
20
3 V s–1
20 V s–1 15 V s–1
1st
500th
1000th
5000th
10,000th
15,000th
20,000th
25,000th
30,000th
10 V s–1 5 V s–1
1 V s–1
2 V s–1
40
20
0
–20
Current (mA cm–2)
Capacitance retention (%)
Pressure (MPa)
Current density (mA cm–2)
Current density (mA cm–2)
Discharge density (mA cm–2)
–40
0 0.2 0.4 0.6
Potential (V) Scan rate (V s–1)
0.8 1.0
c
Fig. 2 Characterizations of HAN@SnO
2
as nanostructured current collectors. a Top-view SEM image of HAN@SnO
2
.bCV curves of HAN@SnO
2
//
HAN@SnO
2
device at different scan rates. cThe discharge current as function of the scan rates based on the CV curves of HAN@SnO
2
//HAN@SnO
2
device at different scan rates. dCycling stability of HAN@SnO
2
//HAN@SnO
2
device tested at a scan rate of 1 V s1.eTop-view SEM image of HAN@SnO
2
electrode after 30,000 CV cycles. fPhotographs of HAN@SnO
2
//HAN@SnO
2
device tested under different mechanical extrusion pressure. gCV proles
of HAN@SnO
2
//HAN@SnO
2
device measured at a scan rate of 500 mV s1under different mechanical extrusion pressure. hTop-view, and icross-
sectional SEM image of HAN@SnO
2
electrode after CV test under a mechanical extrusion pressure of 10 MPa, respectively. Scale bar: 500 nm (a,e); 2 μm
(h); 500 nm (i).
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6
4NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
of the HAN, HAN@SnO
2
electrodes are endowed with high
compression strength to sufciently withstand the mechanical
extrusion during the device assembly without collapse or
deformation of the nanoporous structure. As a result,
HAN@SnO
2
electrodes could remain their initial integrated
structure with vertically aligned nanoporous structure and nally,
the HAN@SnO
2
//HAN@SnO
2
device still works very well
without deteriorating the capacitive behavior even under an
applied mechanical extrusion pressure up to 10 MPa (Fig. 2f, g).
The constant CV proles indicate that there is nearly no change
in the ion-accessible surface area and the ion transport resistance
under the continuous mechanical extrusion. Furthermore, the
SEM images (Fig. 2h, i) reveal that the HAN@SnO
2
electrodes
still remain the original nanoporous features after the electro-
chemical test under the continuous mechanical extrusion. Note
that the residues in Fig. 2h, i is the residual glass bers from a
glass microber lter that was used as separator, and partials of
the glass microber lter were damaged and fallen off during the
disassembly of the HAN@SnO
2
//HAN@SnO
2
device. This result
further reveals the excellent mechanical stability of the HAN that
can efciently sustain the mechanical extrusion pressure during
the device assembly process to maintain the integrality of the
nanoelectrodes. Overall, the excellent electrochemical perfor-
mance and the superior structural stability verify the great
capability of HAN@SnO
2
as desirable nanostructured current
collectors to design nanoelectrodes for batteries and
supercapacitors.
Assembly and electrochemical performance of symmetric
MSCs. Following the electrochemical characterization of
HAN@SnO
2
as nanostructured current collectors, a layer of
pseudocapacitive materials was further electrochemically depos-
ited with HAN@SnO
2
as nanostructured current collectors to
produce nanoelectrodes for MSCs, which are magnesium oxide
(MnO
2
) coated HAN@SnO
2
(HAN@SnO
2
@MnO
2
) and poly-
pyrrole (PPy) coated HAN@SnO
2
(HAN@SnO
2
@PPy) nanoelec-
trodes, respectively. The pore depth of HAN in both two
nanoelectrodes was 25 μm, and the mass loading of MnO
2
and
PPy was 0.65 and 1.32 mg cm2, respectively. The detailed cross-
sectional SEM images of different sections (Supplementary Fig. 3)
and the corresponding energy dispersive X-ray spectroscopy
(EDX) elemental mapping results (Supplementary Figs. 45)
clarify that the cell wall of HAN@SnO
2
have been coated with PPy
and MnO
2
, respectively. Thereafter, MSCs with symmetric device
conguration were assembled through stacking two same HAN-
based nanoelectrodes separated by a glass microber lter and
with 1.0 M Na
2
SO
4
aqueous solution as electrolyte. Note that the
footprint of all stacked MSCs is 0.5 cm2
,
which is equal to that of
single electrode. Figure 3a is the CV curves of the assembled
HAN@SnO
2
@MnO
2
//HAN@SnO
2
@MnO
2
MSCs over a wide
range of scan rates from 2 to 500 mV s1in a potential range of
00.8 V. Obviously, the CV curves at different scan rates exhibit a
symmetrically rectangular shape, indicating an ideally capacitive
behavior and fast chargedischarge characteristics of MSCs. The
galvanostatic chargedischarge (GCD) curves of HAN@S-
nO
2
@MnO
2
//HAN@SnO
2
@MnO
2
MSCs at different current
densities are shown in Fig. 3b. These charging curves are very
symmetric to the discharge counterparts, again indicating the
excellent capacitive behaviors in MSCs. When depositing MnO
2
,
the layer thickness of MnO
2
was kept the same in all HAN@S-
nO
2
@MnO
2
electrodes, thus electrodes with deeper HAN have
higher MnO
2
mass loading. Figure 3c illustrates the device areal
capacitance of MSCs with different pore depth of original HAN
and MnO
2
mass loading as a function of scan rates. The device
capacitance depends strongly on the pore depth of original HAN
and the subsequent changes of MnO
2
mass loading of HAN@S-
nO
2
@MnO
2
electrodes. For HAN@SnO
2
@MnO
2
electrodes con-
sisting of 25-μm-pore-deep HAN, the MnO
2
mass loading was
0.65 mg cm2, consequently resulting in highest device capaci-
tance of 137 mF cm2at a scan rate of 2 mV s1(121 mF cm2at
a current density of 0.2 mA cm2, Supplementary Fig. 6). Bene-
ting from the vertically aligned nanoporous structure features,
these symmetric MSCs have good rate performance, i.e., the
capacitance of MSCs based on HAN@SnO
2
@MnO
2
electrodes
with 25-μm-deep HAN remains to be 47 mF cm2when the scan
rate increasing from 10 to 2,000 mV s1. Additionally, the
HAN@SnO
2
@PPy//HAN@SnO
2
@PPy MSCs have the similar
electrochemical performance to that of HAN@SnO
2
@MnO
2
//
HAN@SnO
2
@MnO
2
MSCs except the different operating poten-
tial window that is from 0.8 to 0 V. Figure 3d, e is the CV and
GCD curves of HAN@SnO
2
@PPy//HAN@SnO
2
@PPy MSCs,
respectively. The pore depth of original HAN in electrodes was 25
μm and the corresponding PPy mass loading was 1.32 mg cm2,
and in this case, the maximum device capacitance of HAN@S-
nO
2
@PPy//HAN@SnO
2
@PPy MSCs reaches 124 mF cm2at a
scan rate of 10 mV s1(158 mF cm2at a current density of 0.2
mA cm2, Supplementary Fig. 6).
By contrast, HAN@SnO
2
//HAN@SnO
2
MSCs with the same
footprint (0.5 cm2) were also assembled and characterized to
further understand the role of SnO
2
layer in HAN@SnO
2
based
pseudocapacitive electrodes. It can be concluded from Fig. 3c, f
that the negligible capacitance (<4 mF cm2) of HAN@SnO
2
//
HAN@SnO
2
MSCs, in which the pore depth of HAN in all
electrodes was 25 μm, suggests the function of SnO
2
layer in
HAN@SnO
2
based pseudocapacitive electrodes mostly for charge
transport rather than charge storage. Furthermore, electrochemi-
cal impendence spectroscopy (EIS) was performed in order to
clearly understand the role of HAN@SnO
2
as nanostructured
current collectors. Figure 4a is the Nyquist plots of HAN@SnO
2
//
HAN@SnO
2
MSCs with different pore depth of original HAN
(i.e., 5, 16 and 25 μm, respectively). All MSCs have very similar
equivalent series resistances (ESR), however, there is a signicant
raise in R
ct
. The gradually increased R
ct
should be the inevitable
result arising from the limited electrical conductivity of SnO
2
,
which is much lower than that of metals. The intrinsic electrical
resistance of SnO
2
leads the total resistance of HAN@SnO
2
//
HAN@SnO
2
MSCs to be progressively increased accompanying
with HAN pore depth increase, and then consequently results in
an increase in R
ct
. With respect to HAN@SnO
2
@MnO
2
//
HAN@SnO
2
@MnO
2
and HAN@SnO
2
@PPy//HAN@SnO
2
@PPy
MSCs (Fig. 4b, c, respectively), all exhibit similar Nyquist curves,
including a semicircle in the high-frequency region and a nearly
vertical line in the low-frequency region. In the high-frequency
region, the diameters of the semicircle for HAN@SnO
2
@MnO
2
//
HAN@SnO
2
@MnO
2
MSCs are bigger than those of HAN@S-
nO
2
@PPy//HAN@SnO
2
@PPy MSCs, indicating a higher R
ct
of
HAN@SnO
2
@MnO
2
//HAN@SnO
2
@MnO
2
MSCs. Additionally,
the higher R
ct
of HAN@SnO
2
@MnO
2
//HAN@SnO
2
@MnO
2
MSCs are also evidenced by the more obvious potential drops
of GCD curves than those of HAN@SnO
2
@PPy//HAN@S-
nO
2
@PPy MSCs (Fig. 3b, e). In addition, both HAN@S-
nO
2
@MnO
2
//HAN@SnO
2
@MnO
2
and HAN@SnO
2
@PPy//
HAN@SnO
2
@PPy MSCs have the similar ESR as those of
HAN@SnO
2
//HAN@SnO
2
MSCs. The similar ESR suggests the
negligible inuence on the series bulk resistance with the use of
HAN@SnO
2
as nanostructured current collectors. In addition, the
smaller ESR in HAN@SnO
2
@PPy//HAN@SnO
2
@PPy MSCs
(Fig. 4c) could be attributed to the higher electrical conductivity
of PPy than that of MnO
2
. Figure 4d summarizes the R
ct
of all
HAN-based MSCs. For HAN@SnO
2
@MnO
2
//HAN@S-
nO
2
@MnO
2
MSCs, the R
ct
are all higher than those of
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6 ARTICLE
NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
HAN@SnO
2
//HAN@SnO
2
MSCs. Besides the increased electrical
resistance in HAN@SnO
2
current collectors, the gradual raise in
R
ct
of HAN@SnO
2
@MnO
2
//HAN@SnO
2
@MnO
2
MSCs should
be due to the low intrinsic conductivity of MnO
2
. The low
conductive MnO
2
layer makes the charge transport in HAN@S-
nO
2
@MnO
2
electrodes mainly rely on the conductive SnO
2
layer,
which results in the exclusive pathway for charge transportation.
On the contrary, the highly conductive PPy layer could not only
store charges but also partially contribute to the charge transport
for improving the charge transport efciency in HAN@S-
nO
2
@PPy electrodes, resulting in lower R
ct
of HAN@S-
nO
2
@PPy//HAN@SnO
2
@PPy MSCs than those in both
HAN@SnO
2
//HAN@SnO
2
and HAN@SnO
2
@MnO
2
//HAN@S-
nO
2
@MnO
2
MSCs. Furthermore, fast ion transport at the
interface between electrode and electrolyte could be evident by
the inconspicuous Warburg region. In the low-frequency region
(Supplementary Fig. 7), all HAN-based MSCs exhibit an almost
vertical line, representing the promising permeability for electro-
lyte inltration and ion diffusion to access the surface of
pseudocapacitive materials, and to create more active sites for
electrochemical reactions to store more charges.
Construction and performance evaluation of asymmetric
MSCs. The areal energy density is directly proportional to the
areal capacitance value and the square of the cell voltage.
Nevertheless, the energy density of supercapacitors is restricted by
the use of pseudocapacitive materials with a narrow potential
window. An asymmetric device conguration affords the
opportunity for the expansion of the operating potential window
of pseudocapacitors, and subsequently accomplishes the
enhanced energy and power densities. Asymmetric MSCs were
therefore assembled by using HAN@SnO
2
@MnO
2
as positive
electrode and HAN@SnO
2
@PPy as negative electrode. The two
electrodes were stacked together with a glass microber lter as
separator and 1.0 M Na
2
SO
4
aqueous solution as electrolyte. As
the aforementioned symmetric MSCs, the footprint of HAN@S-
nO
2
@MnO
2
//HAN@SnO
2
@PPy MSCs was also 0.5 cm2. Note
that a charge balance between the two electrodes was accom-
plished by controlling the deposition time of MnO
2
at the positive
electrode and the thickness of the PPy lm at the negative elec-
trode. Figure 5a outlines the working potential windows of each
symmetric MSCs individually, 0.80 V for HAN@SnO
2
@PPy//
HAN@SnO
2
@PPy MSCs and 00.8 V for HAN@SnO
2
@MnO
2
//
HAN@SnO
2
@MnO
2
MSCs, so that the asymmetric MSCs by
integrating HAN@SnO
2
@PPy and HAN@SnO
2
@MnO
2
into
single MSCs will result in an increased operating cell voltage up to
1.6 V. As shown in Fig. 5b, the asymmetric MSCs with 16-μm-
pore-deep HAN in all electrodes work efciently in the range
from 0.8 to 1.6 V. Figure 5c, d is the corresponding CV and GCD
curves of asymmetric MSCs in a potential range of 01.6 V,
respectively. The nearly rectangular CV curves and the highly
triangular chargedischarge proles demonstrate an ideally
capacitive behavior and the fast chargedischarge characteristics
of the asymmetric MSCs. Particularly, the CV curves retain their
rectangular shape without apparent distortions when increasing
scan rates up to a high rate of 1000 mV s1, indicating the
extraordinary high-rate performance of asymmetric MSCs. And a
highest device capacitance of 147 mF cm2has been achieved at a
scan rate of 10 mV s1and still remain 80 mF cm2at the scan
rate of 1000 mV s1(Supplementary Fig. 6). The maximum
capacity of this device could reach 186 mF cm2at a current
density of 0.2 mA cm2(Supplementary Fig. 8), and is one of the
highest values among the reported MSCs (Supplementary
Table 1). Moreover, the asymmetric MSCs have an excellent
cyclic capability of 87% of its original capacity after continued
30,000 charge/discharge cycles tested at a current density of 20
mA cm2and with nearly 100% Coulombic efciency (Fig. 5f).
The slight performance degradation should be mainly due to
MnO
2
, as evidenced by the obvious morphological changes after
cycling (Supplementary Fig. 9).
60
a
def
bc
0.8
0.6
0.4
0.2
0.0
–0.8
–0.6
–0.4
–0.2
0.0
0 100 200 300
Time (s) Scan rate (mV s–1)
400 500
0 100 200 300
Time (s)
400 500
160
120
200
150
100
50
0
80
40
0
0 500 1000 1500 2000
Scan rate (mV s–1)
0 500 1000 1500 2000
500 mV s–1 200 mV s–1 100 mV s–1
0.5 mA cm–2
25 μm@0.65 mg cm–2
25 μm@1.32 mg cm–2
16 μm@0.83 mg cm–2
5 μm@0.21 mg cm–2
25 μm@0 mg cm–2
16 μm@0.43 mg cm–2
25 μm@0 mg cm–2
MnO2 mass loading
PPy mass loading
Pore depth
Pore depth
5 μm@0.11 mg cm–2
1 mA cm–2
2 mA cm–2
5 mA cm–2
10 mA cm–2
0.5 mA cm–2
1.0 mA cm–2
2.0 mA cm–2
5.0 mA cm–2
10 mA cm–2
10 mV s–1
100 mV s–1
10 mV s–1
20 mV s–1
2 mV s–1
200 mV s–1
20 mV s–1
2 mV s–1
50 mV s–1
5 mV s–1
500 mV s–1
50 mV s–1
5 mV s–1
40
20
Current (mA cm–2)Current (mA cm–2)
Potential (V)
Potential (V)
Areal capacitance (mF cm–2)Areal capacitance (mF cm–2)
–20
–40
–40
–20
0
20
40
60
80
–60
0 0.2 0.4
Potential (V)
Potential (V)
0.6 0.8
0.0–0.2–0.4–0.6–0.8
0
Fig. 3 Electrochemical performance of symmetric MSCs based on HAN-based nanoelectrodes. a CV curves at different scan rates, bGCD proles at
different current densities, and cdevice areal capacitance as a function of scan rates of HAN@SnO
2
@MnO
2
//HAN@SnO
2
@MnO
2
MSCs, respectively. d
CV curves at different scan rates, eGCD proles at different current densities, and fdevice areal capacitance as a function of scan rates of
HAN@SnO
2
@PPy//HAN@SnO
2
@PPy MSCs, respectively.
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6
6NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
When further replacing aqueous electrolyte with the ionic
liquid electrolyte (1-ethyl-3-methylimidazolium bis(triuoro-
methylsulfonyl)imide, EMIM-TFSI), the asymmetric MSCs can
work efciently in an enlarged potential range of 03.0 V (Fig. 5g,
h). The distorted CV and GCD curves arise from the high
viscosity and large charge transfer resistance of the electrolytes
because of the large ion size of the ionic liquids22,23. The device
capacitance reaches 128 mF cm2at a current density of 0.5 mA
cm2(or 162 mF cm2at a scan rate of 10 mV s1). When the
current density is increased from 0.5 to 5 mA cm2, the
capacitance drops to 93 mF cm2(72% retention) and it recovers
to 123 mF cm2(96% retention) with the current density again
being decreased to 0.5 mA cm2(Fig. 5i), revealing the excellent
rate performance of the asymmetric MSCs. In addition, the
asymmetric MSCs with ionic liquid electrolyte still exhibit
remarkable cycling stability with 82.5% of initial capacitance at
20 mA cm2over a potential window of 03.0 V withstanding
continued 10,000 cycles with high Coulombic efciency of nearly
100% and slight decay in CV curves at 100 mV s1before and
after cycling (Supplementary Fig. 10). Figure 6a presents the areal
energy performance metrics of one representative HAN@S-
nO
2
@MnO
2
//HAN@SnO
2
@PPy asymmetric MSCs with EMIM-
TFSI ionic liquid electrolyte. The maximum capacitance of the
MSCs reaches 128 mF cm2at a current density of 0.5 mA cm2,
and the peak energy and power densities are 160 μWh cm2and
40 mW cm2, respectively. Remarkably, the energy storage
performance of the asymmetric MSCs with HAN-based nanoe-
lectrodes is among the best comprehensive performance of the
reported stacked MSCs (Fig. 6b)21,2434. In particular, the peak
energy density of the stacked asymmetric MSCs in this work is
roughly fourfold that of the carbide-derived-carbons (CDC)
based stacked MSCs (~40 μWh cm2) but with a similar peak
power density24. Additionally, the areal energy density of the
asymmetric MSCs with EMIM-TFSI electrolyte is even compar-
able with that of some state-of-the-art 3D micro-batteries but
with much higher areal power density35,36.
Discussion
The insufcient energy of MSCs is their crucial aw compared
with micro-batteries and remains a major challenge to overcome.
This work signies the capability of the HAN as a promising
nanostructuring platform to assemble pseudocapacitive materials
toward rationally designing nanoelectrodes for MSCs with high-
energy performance. In comparison with the conventional HMs,
the HAN as a kind of microminiaturized HMs has the highest cell
density, lower relative density, higher surface area enhancement
factor, and excellent compressive performance. With such robust
HAN as the keel, the as-prepared nanoelectrodes could possess
vertically aligned and robustly stable nanoporous structure. And
in contrast to nanoelectrodes design based on nanowires and
nanotubes, there is no limit about the aspect ratio to sustain such
vertically aligned nanoporous structure. Besides high surface area
to provide abundant electroactive sites for surface Faradaic
reactions, the vertically aligned nanoporous structure affords the
smooth channels to favor the permeability for electrolyte
10
a
c
b
d
10
0.11 mg cm–2
0.21 mg cm–2 0.83 mg cm–2
1.32 mg cm–2
0.43 mg cm–2
0.65 mg cm–2
88
66
–Z′′ (Ω)
–Z′′ (Ω)
–Z′′ (Ω)
Z (Ω)
Z (Ω)
44
22
0
0
1
2
3
4
5
0
8
6
4
Rct (Ω)
2
0
HAN@SnO2
HAN@SnO2@MnO2
HAN@SnO2@PPy
5 μm5 μm
5 μm
16 μm16 μm
16 μm
25 μm
5 μm
16 μm
25 μm
25 μm
25 μm
0
2
3
4
5
1
0
2
4
6
8
10
Z (Ω)
0
2
4
6
8
10
Fig. 4 Electrochemical impedance properties comparison. Nyquist plots of aHAN@SnO
2
//HAN@SnO
2
,bHAN@SnO
2
@MnO
2
//HAN@SnO
2
@MnO
2
,
and cHAN@SnO
2
@PPy//HAN@SnO
2
@PPy MSCs with different pore depths of HAN. dComparison of the charge transport resistance (R
ct
) of different
HAN-based MSCs.
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6 ARTICLE
NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
a
d
ghi
ef
bc
15
10
5
0
–5
–10 –80
–40
0
40
80
120
0 0.4 0.8 1.2 1.6
3
Current (mA cm–2)
Current (mA cm–2)
Current (mA cm–2)
Current (mA cm–2)
0
–3
–6
–0.8
1.6
1.2
0.8
0.4
0
0
20
40
–20
–40
0 200 400
Time (s)
Potential (V)
Scan rate (m V s–1)
600
160
120
80
40
0
0
2
3
1
0 200 400 600 800
800600400
Time (s)
20000123
1000
100
150
100
50
0
0 1020304050
100
80
60
Coulombic efficiency (%)
40
20
0
80 1.6
1.2
0.8
0.4
0.0
0.00 0.01 0.02 73.68
Time (h)
73.69 73.70
60
40
20
0
0 5000 10,000 15,000
Cycle number
Cycle number
20,000 25,000 30,000
–0.4 0
Potential (V)
Potential (V)
Potential (V)
Potential (V)
Areal capacitance (mF cm–2)
Areal capacitance (mF cm–2)Capacitance retention (%)
Potential (V)
0 0.4 0.8 1.2 1.6
Potential (V)
0.4
0–0.8 V
–0.8–0 V
HAN@SnO2@MnO2
HAN@SnO2@PPy
0.8
@ 100 mV s–1
@ 20 mA cm–2
0.5 mA cm–2
1 mA cm–2
2 mA cm–2
5 mA cm–2
10 mA cm–2
15 mA cm–2
25 mA cm–2
5 mA cm–2
0.5 mA cm–2
@ 100 mV s–1
1000 mV s–1
200 mV s–1 150 mV s–1
20 mV s–1 10 mV s–1
100 mV s–1
50 mV s–1
0.5 mA cm–2 25 μm
16 μm
5 μm
1.0 mA cm–2
1 mA cm–2
2 mA cm–2
5 mA cm–2
10 mA cm–2
15 mA cm–2
20 mA cm–2
2.0 mA cm–2
5.0 mA cm–2
10 mA cm–2
500 mV s–1 200 mV s–1
20 mV s–1
50 mV s–1
5 mV s–1
100 mV s–1
10 mV s–1
1.6 V 1.4 V 1.2 V 1.0 V
0.8 V
Fig. 5 Electrochemical performance of HAN@SnO
2
@MnO
2
//HAN@SnO
2
@PPy asymmetric MSCs. a Typical CV curves of HAN@SnO
2
@MnO
2
and
HAN@SnO
2
@PPy based symmetric MSCs, respectively, with 1.0 M Na
2
SO
4
electrolyte. bCV curves within different potential ranges. cCV curves at
different scan rates. dGCD proles at different current densities. eDevice areal capacitance as a function of current densities. fCycling stability test at a
scan rate of 20 mA cm2.gCV curves at different scan rates, hGCD proles at different current densities, and irate performance of asymmetric MSCs
with EMIM-TFSI electrolyte, respectively.
120
ab
1020.36 s
10 h
This work, MnO2//PPy
(in Na2SO4)
This work, MnO2//PPy
(in EMIMTFSI)
This work, MnO2//MnO2
LWG, ref. 28
rGO MSCs, ref. 29
PG-MSCs, ref. 30
MPG-MSCs, ref. 32
SG-MSCs, ref. 31
G-CNTCs, ref. 21
Textile MSCs, ref. 34
OLC, ref. 25
LSG-MSCs,
ref. 27
CDC, ref. 24
LSG/MnO2, ref. 26
This work
PPy//PPy
1 h 360 s 36 s 3.6 s
36 ms
3.6 ms
0.36 ms
36 μs
100
10–2
Energy density (μ Wh cm–2)
Power density (mW cm–2)
10–4
10–3 10–1 101103
Capacitance (mF cm
–2
)
Power density (mW cm
–2
)
Energy density (μwh cm
–2
)
80
40
0
40
40
80
120
160
30
20
10
0
Fig. 6 Energy performance of HAN@SnO
2
@MnO
2
//HAN@SnO
2
@PPy asymmetric MSCs. a Areal performance metrics of HAN@SnO
2
@MnO
2
//
HAN@SnO
2
@PPy asymmetric MSCs with EMIM-TFSI ionic liquid electrolyte. bRagone plots of MSCs with HAN-based nanoelectrodes compared with
some reported MSCs21,2434.
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6
8NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
inltration and ion diffusion to access the surface of pseudoca-
pacitive materials, guaranteeing a high utilization efciency of the
pseudocapacitive materials. Taken together, the HAN-based
nanoelectrodes design synergizes the effects of both effective
ion migration and high electroactive surface area, thus enabling
high and reversible capacitive behavior even at high
chargedischarge rates. In the current HAN-based nanoelec-
trodes design, four strategies have been combined together into
single MSCs aiming to achieve high-energy storage capability: (i)
the utilization of pseudocapacitive materials with high specic
capacitance in both positive and negative electrodes; (ii) the
adoption of asymmetric device conguration to extend the
operating potential window of MSCs; (iii) the application of ionic
liquid electrolyte to further widen the operating potential window
of MSCs; and more importantly, (iv) the prominent nanoelec-
trode architecture design based on HAN for both positive and
negative electrodes. The combination of pseudocapacitive mate-
rials with unique nanoelectrodes architecture results in the
increased areal capacitance with reduced footprint, meanwhile
the adoption of asymmetric device conguration coupled with
ionic liquid electrolyte helps to overcome the narrow operating
potential window of pseudocapacitive materials. As a result, the
asymmetric MSCs constructed with the HAN-based nanoelec-
trodes exhibit one of the most remarkable comprehensive areal
device performance metrics.
We emphasize that nanoelectrode design with the robust HAN
as the keel to assemble electroactive materials inherits the unique
structural characteristics of the HMs and can be manipulated to
achieve desired application requirements for electrochemical
devices beyond MSCs. Our work opens up the ample opportu-
nities for further expanding the application range of the HMs and
also provides a paradigm about rationally designing nanos-
tructures for various functional devices with new features and
high performance.
Methods
Preparation of HAN. The HAN was fabricated through a nanoindentation-
anodization process, followed by a precisely controlled two-step chemical etching
process. Firstly, a hexagonal patterns was generated onto the surface of aluminum foil
by nanoindentation, then the aluminium foil was anodized in a 0.4 M phosphoric acid
(H
3
PO
4
) aqueous solution at 160 V for different time. The temperature ofthe solution
and the aluminium foil was kept at 10°C during the anodization process. After
anodization, the unanodized aluminum was etched by a mixture aqueous solution of
CuCl
2
(85 wt%) and HCl (15wt%). Thereafter a two-step chemical etching process
was applied to obtain the HAN: (1) etching with a 5wt% H
3
PO
4
solution at 60 °C to
achieve an open-end membrane; (2) further etching with a 5 wt% H
3
PO
4
solution at
30 °C to nally obtain HAN. The surface area enhancement factor (A)forHANwas
calculated by Eq. 1:
A¼
6
ffiffi
3
p´d´h
s2´sin θð1Þ
where dand hare corresponding to the cell diameter and depth of the cell, respec-
tively, sis the intercell distances, measured between the centers of the cells, and θis
the angle of the pattern in which the cells are positioned. For the typical HAN
structure with cell depth of 25 μm, the specic values are d=384 nm, h=25 μm, s=
400 nm and θ=60°, respectively. Accordingly, the Ais calculated to be 240.
Fabrication of HAN-based nanoelectrodes. The HAN@SnO
2
substrate was
produced by conformally depositing SnO
2
on the surface of the as-prepared HAN
at 250 °C using an ALD system (Picosun, SUNA LE R-150). The precursors were tin
(IV) chloride (SnCl
4
) for tin and ultrapure water for oxygen, respectively, and high-
purity nitrogen was the carrier and purging gas. Specically, each ALD cycle
consisted of a 0.2 s pulse of SnCl
4
and a 4 s purge of N
2
, followed by a 1 s pulse of
H
2
O and an 8 s purge of N
2
, and this procedure was repeated 1500 times. The
growth rate of SnO
2
is estimated to be about 0.16 Å per cycle. With the as-obtained
HAN@SnO
2
as the working electrodes, the HAN-based nanoelectrodes were fab-
ricated through electrochemical deposition method. The HAN@SnO
2
@PPy elec-
trode was prepared by electrochemical polymerization of PPy onto HAN@SnO
2
substrate. The plating solution consisted of 0.1 M pyrrole monomer (98%) and 0.2
M oxalic acid, and the applied potential was 0.8 V (vs. Ag/AgCl)37. Likewise, the
HAN@SnO
2
@MnO
2
electrode was fabricated by electrochemical deposition of
MnO
2
onto the HAN@SnO
2
substrate by using an electrolyte consisting of 50 mM
manganese acetate and 100 mM sodium acetate with a constant current density of
1mAcm
2. The active mass of the electrodes were determined according to
Faradays law and 100% charge efciency was assumed, according to Eq. 2:
m¼QM
zF ð2Þ
where Qis the charge passed during the electrochemical deposition process of
active materials (i.e., PPy/MnO
2
). And it is worth noting that the charge is precisely
adjust here in order to achieve the same layer thickness of active materials for
HAN@SnO
2
with different cell depth. Mis the molar mass of the active electrode
material (M
MnO2
=86.9 g mol1;M
py
=65.09 g mol1), Fthe Faraday constant
and zthe number of transferred electrons per active electrode atom (z=2 for
MnO
2
and PPy deposition).
Apparatus for characterizations. The surface morphologies and microstructures
of the electrodes were characterized using scanning electron microscopy (ZEISS
AURIGA) equipped with an EDX detector. Cyclic voltammetry (CV), galvanostatic
chargedischarge (GCD), electrochemical impedance spectroscopy (EIS) mea-
surements, and cycling performance measurements were conducted with a
Potentiostat (BioLogic, VSP). Nanoindentation was conducted with the Tri-
boindenter (Hysitron) equipped with a Berkovich diamond indenter with an
approximately 600-nm-radius.
Electrochemical measurements and calculation of the electrochemical per-
formances. All the electrochemical measurements were carried outin a two-electrode
conguration with 1.0 M Na
2
SO
4
aqueous electrolyte and a glass microber lter
(Whatman, GF/B) as separator as well as 1-ethyl-3-methylimidazolium bis(tri-
uoromethylsulfonyl)imide (EMIM-TFSI) ionic liquid electrolyte (1M in acetoni-
trile). The CV curves were collected with various scan rates of 5 to 2000 mVs1,and
the GCD curves were obtained at current densities of 0.2, 0.5, 1, 2, 5, 10, 15, 20 and
25 mA cm2. The EIS measurement was performed with a frequency range from
100 MHz to 10 mHz and a 5 mV AC amplitude.
The capacitance of the device C
GCD
(mF cm2) is determined from the GCD
proles based on Eq. 3:
CGCD ¼I´Δt
A´ΔVð3Þ
where Δt (s) is the discharge time, I(mA) is the discharge current, A(cm2) is total
footprint area of device, and ΔV (V) is the voltage window.
The device capacitance C
CV
(mF cm2) can also be calculated from CV curves
based on Eq. 4:
CCV ¼RIdV
υ´ΔV´Að4Þ
where I(A) is the response current and ν(V s1) is the potential scan rate.
With GCD plots, the energy density (E, unit: μWh cm2) and power density (P,
unit: mW cm2) of the device were calculated by Eqs. 5and 6:
E¼0:5´C´ΔV2ð5Þ
P¼E
Δtð6Þ
where C(mF cm2) is the areal capacitance of the device, Δt (s) is the
discharge time.
Date availability
The data that support the ndings of this study are available from the corresponding
authors on reasonable request.
Received: 22 April 2019; Accepted: 13 December 2019;
References
1. Gibson, L. J. & Ashby, M. F. Cellular Solids: Structure and Properties.
(Cambridge Univ. Press, 1999).
2. Kreutzer, M. T., Kapteijn, F., Moulijn, J. A. & Heiszwolf, J. J. Multiphase
monolith reactors: chemical reaction engineering of segmented ow in
microchannels. Chem. Eng. Sci. 60, 58955916 (2005).
3. Cybulski, A. & Moulijn, J. A. Monoliths in heterogeneous catalysis. Catal. Rev.
Sci. Eng. 36, 179270 (1994).
4. Avila, P., Montes, M. & Miró, E. E. Monolithic reactors for environmental
applications: a review on preparation technologies. Chem. Eng. J. 109,1136
(2005).
5. Williams, J. L. Monolith structures, materials, properties and uses. Catal.
Today 69,39 (2001).
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6 ARTICLE
NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications 9
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6. Guo, Y. et al. Robust 3-D congurated metal oxide nano-array based
monolithic catalysts with ultrahigh materials usage efciency and catalytic
performance tunability. Nano Energy 2, 873881 (2013).
7. Ren, Z. et al. Monolithically integrated spinel M
x
Co
3x
O
4
(M =Co, Ni, Zn)
nanoarray catalysts: scalable synthesis and cation manipulation for tunable
lowtemperature CH
4
and CO oxidation. Angew. Chem. Int. Ed. 53,
72237227 (2014).
8. Zhao, H., Zhou, M., Wen, L. & Lei, Y. Template-directed construction of
nanostructure arrays for highly-efcient energy storage and conversion. Nano
Energy 13, 790813 (2015).
9. Aricò, A. S., Bruce, P., Scrosati, B., Tarascon, J.-M. & Van Schalkwijk, W.
Nanostructured materials for advanced energy conversion and storage devices.
Nat. Mater. 4, 366377 (2005).
10. Manthiram, A., Murugan, A. V., Sarkar, A. & Muraliganth, T. Nanostructured
electrode materials for electrochemical energy storage and conversion. Energy
Environ. Sci. 1, 621638 (2008).
11. Li, Y., Fu, Z. Y. & Su, B. L. Hierarchically structured porous materials for
energy conversion and storage. Adv. Funct. Mater. 22, 46344667 (2012).
12. Thompson, G. & Wood, G. Porous anodic lm formation on aluminium.
Nature 290, 230 (1981).
13. Li, G. et al. Synthesis of biomorphic zeolite honeycomb monoliths with 16000
cells per square inch. J. Mater. Chem. 19, 83728377 (2009).
14. Cannarella, J. & Arnold, C. B. Stress evolution and capacity fade in
constrained lithium-ion pouch cells. J. Power Sources 245, 745751 (2014).
15. Fukushima, M. & Yoshizawa, Y.-i Fabrication of highly porous honeycomb-
shaped mullitezirconia insulators by gelation freezing. Adv. Powder Technol.
27, 908913 (2016).
16. Yu, Z.-L. et al. Bioinspired polymeric woods. Sci. Adv. 4, eaat7223 (2018).
17. Zheng, X. et al. Ultralight, ultrastiff mechanical metamaterials. Science 344,
13731377 (2014).
18. Bauer, J., Hengsbach, S., Tesari, I., Schwaiger, R. & Kraft, O. High-strength
cellular ceramic composites with 3D microarchitecture. Proc. Natl Acad. Sci.
USA 111, 24532458 (2014).
19. Schaedler, T. A. et al. Ultralight metallic microlattices. Science 334, 962965 (2011).
20. Li, T. et al. Anisotropic, lightweight, strong, and super thermally insulating
nanowood with naturally aligned nanocellulose. Sci. Adv. 4, eaar3724 (2018).
21. Lin, J. et al. 3-dimensional graphene carbon nanotube carpet-based
microsupercapacitors with high electrochemical performance. Nano Lett. 13,
7278 (2012).
22. Kim, T., Jung, G., Yoo, S., Suh, K. S. & Ruoff, R. S. Activated graphene-based
carbons as supercapacitor electrodes with macro-and mesopores. ACS Nano 7,
68996905 (2013).
23. Zhang, F. et al. A high-performance supercapacitor-battery hybrid energy
storage device based on graphene-enhanced electrode materials with ultrahigh
energy density. Energy Environ. Sci. 6, 16231632 (2013).
24. Huang, P. et al. On-chip and freestanding elastic carbon lms for micro-
supercapacitors. Science 351, 691695 (2016).
25. Pech, D. et al. Ultrahigh-power micrometre-sized supercapacitors based on
onion-like carbon. Nat. Nanotechnol. 5, 651654 (2010).
26. El-Kady, M. F. et al. Engineering three-dimensional hybrid supercapacitors
and microsupercapacitors for high-performance integrated energy storage.
Proc. Natl Acad. Sci. USA 112, 42334238 (2015).
27. El-Kady, M. F. & Kaner, R. B. Scalable fabrication of high-power graphene
micro-supercapacitors for exible and on-chip energy storage. Nat. Commun.
4, 1475 (2013).
28. Gao, W. et al. Direct laser writing of micro-supercapacitors on hydrated
graphite oxide lms. Nat. Nanotechnol. 6, 496500 (2011).
29. Niu, Z. et al. Allsolid-state exible ultrathin microsupercapacitors based on
graphene. Adv. Mater. 25, 40354042 (2013).
30. Xiao, H. et al. One-step device fabrication of phosphorene and graphene
interdigital micro-supercapacitors with high energy density. ACS Nano 11,
72847292 (2017).
31. Wu, Z.-S. et al. Bottom-up fabrication of sulfur-doped graphene lms derived
from sulfur-annulated nanographene for ultrahigh volumetric capacitance
micro-supercapacitors. J. Am. Chem. Soc. 139, 45064512 (2017).
32. Wu, Z. S., Parvez, K., Feng, X. & Müllen, K. Graphene-based in-plane micro-
supercapacitors with high power and energy densities. Nat. Commun. 4, 2487
(2013).
33. Shen, K., Ding, J. & Yang, S. 3D printing quasisolidstate asymmetric micro
supercapacitors with ultrahigh areal energy density. Adv. Energy Mater. 8,
1800408 (2018).
34. Pu, X. et al. Wearable textilebased inplane microsupercapacitors. Adv.
Energy Mater. 6, 1601254 (2016).
35. Pikul, J. H., Zhang, H. G., Cho, J., Braun, P. V. & King, W. P. High-power
lithium ion microbatteries from interdigitated three-dimensional
bicontinuous nanoporous electrodes. Nat. Commun. 4, 1732 (2013).
36. Ning, H. et al. Holographic patterning of high-performance on-chip 3D
lithium-ion microbatteries. Proc. Natl Acad. Sci. USA 112,65736578
(2015).
37. Grote, F. & Lei, Y. A complete three-dimensionally nanostructured
asymmetric supercapacitor with high operating voltage window based on PPy
and MnO
2
.Nano Energy 10,6370 (2014).
Acknowledgements
The authors gratefully acknowledge the support from German Research Foundation
(DFG: LE 2249/4-1 and LE 2249/5-1), as well as nancial support for Long Liu from the
China Scholarship Council (CSC).
Author contributions
Zh. L and Lo. L contributed equally to this work. Y.L. supervised the project. The concept
was conceived by H.Z. and Y.L. Zh. L, Lo. L and H.Z. were involved in the fabrication and
related characterizations. Zh. L, Lo. L, S.C. and Le. L performed the electrochemical
experiments and data analysis under the guidance of H.Z., F.L., Zh. L, J.K. and Y.L. H.Z.
and Y.L. wrote the paper with the assistance of F.L., Y.Z. and J.K.
Competing interests
The authors declare no competing interests.
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41467-
019-14170-6.
Correspondence and requests for materials should be addressed to H.Z., F.L., Z.L. or Y.L.
Peer review information Nature Communications thanks the anonymous reviewer(s) for
their contribution to the peer review of this work. Peer reviewer reports are available.
Reprints and permission information is available at http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party
material in this article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not included in the
articles Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
© The Author(s) 2020
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-14170-6
10 NATURE COMMUNICATIONS | (2020) 11:299 | https://doi.org/10.1038/s41467-019-14170-6 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... However, the development of supercapacitors using cost-effective and eco-friendly materials presents a persistent challenge. Supercapacitor materials are broadly classified into electrochemical double-layer capacitors (EDLCs), which store energy through non-faradaic charge separation, and pseudocapacitive materials, which rely on fast, reversible redox reactions for energy storage [7][8][9][10]. A promising strategy to achieve superior performance involves hybrid capacitors, which combine carbon-based materials with metal oxides and conducting polymers [11]. ...
Article
Full-text available
This study presents the successful synthesis and characterization of polyaniline (PANI), PANI/reduced graphene oxide PANI/rGO (PR), and PANI/rGO/ZnO (PRZ) nanocomposites as electrode materials for supercapacitors. Employing electrospinning and electrospraying techniques, we developed nanofibrous composites with enhanced structural and electrochemical properties. The addition of rGO and ZnO in the PRZ composite significantly improved specific capacitance, stability, and charge‐transfer efficiency. Electrochemical analyses, including cyclic voltammetry (CV), galvanostatic charge–discharge (GCD), and electrochemical impedance spectroscopy (EIS), revealed a peak specific capacitance of 845 F g⁻¹ at 0.5 A g⁻¹ for PRZ, outperforming PR (395 F g⁻¹), and PANI (140 F g⁻¹). These enhancements are attributed to the synergistic effects of carbon‐based and pseudocapacitive components, resulting in higher conductivity, improved redox activity, and reduced internal resistance. Additionally, the PRZ composite exhibited excellent cyclic stability, retaining 89% of its capacitance over 5000 cycles, underscoring its durability and suitability for long‐term energy storage applications.
... Cyclic voltammetry (CV) and galvanostatic charge−discharge profiles were recorded at the potential window between 0 and 1 V to evaluate the electrochemical performance of the EC part in mp-SC. The nearly rectangular CV curves (Fig. 2d) indicate the typical double-layer capacitive behavior of the rGO microelectrodes 28,29 . The area-specific capacitance of the EC part according to the CV profiles is shown in Fig. 2e, which shows the areal capacitance of 2.21 mF cm -2 at a current density of 10 mA cm -2 , comparable to values of the state-of-the-art carbon-based supercapacitors 28 . ...
Article
Full-text available
Supercapacitor is highly demanded in emerging portable electronics, however, which faces frequent charging and inevitable rapid self-discharging of huge inconvenient. Here, we present a flexible moisture-powered supercapacitor (mp-SC) that capable of spontaneously moisture-enabled self-charging and persistently voltage stabilizing. Based on the synergy effect of moisture-induced ions diffusion of inner polyelectrolyte-based moist-electric generator and charges storage ability of inner graphene electrochemical capacitor, this mp-SC demonstrates the self-charged high areal capacitance of 138.3 mF cm⁻² and ~96.6% voltage maintenance for 120 h. In addition, a large-scale flexible device of 72 mp-SC units connected in series achieves a self-charged 60 V voltage in air, efficiently powering various commercial electronics in practical applications. This work will provide insight into the design self-powered and ultra-long term stable supercapacitors and other energy storage devices.
Article
In the last decade, the diversity of high-entropy materials (HEMs) has increased sharply, including due to the expansion of research into the field of amorphous, nano- and heterostructures. Interest in nanoscale HEMs is primarily associated with their potential application in various fields, such as renewable and green energy, catalysis, hydrogen storage, surface protection and others. The development of nanotechnology has made it possible to develop an innovative design of nanoscale HEMs with fundamentally new structures with unique physical and chemical properties. Problems of controlled synthesis with precisely specified parameters of chemical composition, microstructure and morphology are solved. At the same time, traditional technologies such as fast pyrolysis, mechanical alloying, magnetron sputtering, electrochemical synthesis, etc. are being modernized. Along with this, innovative synthesis technologies have appeared, such as carbothermic shock, the method of controlled hydrogen spillover. The review discusses various methods for the synthesis of nanoscale HEMs that have been developed in the last few 6–7 years for various applications. Some of them are modernization of traditional methods for producing HEM or nano-sized materials, while another group of techniques represents innovative solutions stimulated and inspired by the HEM phenomenon.
Article
Full-text available
The intensifying energy crisis has made it urgent to develop robust and reliable next‐generation energy systems. Except for conventional large‐scale energy sources, the imperceptible and randomly distributed energy embedded in daily life awaits comprehensive exploration and utilization. Harnessing the latent energy has the potential to facilitate the further evolution of soft energy systems. Compared with rigid energy devices, flexible energy devices are more convenient and suitable for harvesting and storing energy from dynamic and complex structures such as human skin. Stretchable conductors that are capable of withstanding strain (≫1%) while sustaining stable conductive pathways are prerequisites for realizing flexible electronic energy devices. Therefore, understanding the characteristics of these conductors and evaluating the feasibility of their fabrication strategies are particularly critical. In this review, various preparation methods for stretchable conductors are carefully classified and analyzed. Furthermore, recent progress in the application of energy harvesting and storage based on these conductors is discussed in detail. Finally, the challenges and promising opportunities in the development of stretchable conductors and integrated flexible energy devices are highlighted, seeking to inspire their future research directions.
Article
Full-text available
Polymer cubosomes (PCs) have well‐defined inverse bicontinuous cubic mesophases formed by amphiphilic block copolymer bilayers. The open hydrophilic channels, large periods, and robust physical properties of PCs are advantageous to many host–guest interactions and yet not fully exploited, especially in the fields of functional nanomaterials. Here, the self‐assembly of poly(ethylene oxide)‐block‐polystyrene block copolymers is systematically investigated and a series of robust PCs is developed via a cosolvent method. Ordered nanoporous metal oxide particles are obtained by selectively filling the hydrophilic channels of PCs via an impregnation strategy, followed by a two‐step thermal treatment. Based on this versatile PC platform, the general synthesis of a library of ordered porous particles with different pore structures 3¯3ˉ\bar{3}3ˉ\bar{3}, tunable large pore size (18–78 nm), high specific surface areas (up to 123.3 m² g⁻¹ for WO3) and diverse framework compositions, such as transition and non‐transition metal oxides, rare earth chloride oxides, perovskite, pyrochlore, and high‐entropy metal oxides is demonstrated. As typical materials obtained via this method, ordered porous WO3 particles have the advantages of open continuous structure and semiconducting properties, thus showing superior gas sensing performances toward hydrogen sulfide.
Article
Full-text available
Woods provide bioinspiration for engineering materials due to their superior mechanical performance. We demonstrate a novel strategy for large-scale fabrication of a family of bioinspired polymeric woods with similar polyphenol matrix materials, wood-like cellular microstructures, and outstanding comprehensive performance by a self-assembly and thermocuring process of traditional resins. In contrast to natural woods, polymeric woods demonstrate comparable mechanical properties (a compressive yield strength of up to 45 MPa), preferable corrosion resistance to acid with no decrease in mechanical properties, and much better thermal insulation (as low as ~21 mW m⁻¹ K⁻¹) and fire retardancy. These bioinspired polymeric woods even stand out from other engineering materials such as cellular ceramic materials and aerogel-like materials in terms of specific strength and thermal insulation properties. The present strategy provides a new possibility for mass production of a series of high-performance biomimetic engineering materials with hierarchical cellular microstructures and remarkable multifunctionality.
Article
Full-text available
There has been a growing interest in thermal management materials due to the prevailing energy challenges and unfulfilled needs for thermal insulation applications. We demonstrate the exceptional thermal management capabilities of a large-scale, hierarchal alignment of cellulose nanofibrils directly fabricated from wood, hereafter referred to as nanowood. Nanowood exhibits anisotropic thermal properties with an extremely low thermal conductivity of 0.03 W/m·K in the transverse direction (perpendicular to the nanofibrils) and approximately two times higher thermal conductivity of 0.06 W/m·K in the axial direction due to the hierarchically aligned nanofibrils within the highly porous backbone. The anisotropy of the thermal conductivity enables efficient thermal dissipation along the axial direction, thereby preventing local overheating on the illuminated side while yielding improved thermal insulation along the backside that cannot be obtained with isotropic thermal insulators. The nanowood also shows a low emissivity of <5% over the solar spectrum with the ability to effectively reflect solar thermal energy. Moreover, the nanowood is lightweight yet strong, owing to the effective bonding between the aligned cellulose nanofibrils with a high compressive strength of 13 MPa in the axial direction and 20 MPa in the transverse direction at 75% strain, which exceeds other thermal insulation materials, such as silica and polymer aerogels, Styrofoam, and wool. The excellent thermal management, abundance, biodegradability, high mechanical strength, low mass density, and manufacturing scalability of the nanowood make this material highly attractive for practical thermal insulation applications.
Article
Full-text available
Heteroatom doping of nanocarbon films can efficiently boost the pseudocapacitance of micro-supercapacitors (MSCs), however, wafer-scale fabrication of sulfur-doped graphene films with a tailored thickness and homogeneous doping for MSCs remains a great challenge. Here we demonstrate the bottom-up fabrication of continuous, uniform, ultrathin sulfur-doped graphene (SG) films, derived from the peripherical tri-sulfur-annulated hexa-peri-hexabenzocoronene (SHBC), for ultrahigh-rate MSCs (SG-MSCs) with landmark volumetric capacitance. The SG film was prepared by thermal annealing of the spray-coated SHBC-based film, with assistance of a thin Au protecting layer, at 800 oC for 30 min. SHBC with twelve phenylthio groups decorated at the periphery is critical as precursor for the formation of the continuous and ultrathin SG film, with a uniform thickness of ~10.0 nm. Notably, the as-produced all-solid-state planar SG-MSCs exhibited a highly stable pseudocapacitive behavior with an volumetric capacitance of ~582 F cm-3 at 10 mV s-1, excellent rate capability with a remarkable capacitance of 8.1 F cm-3 even at an ultrahigh rate of 2000 V s-1, ultrafast frequency response with a short time constant of 0.26 ms, and ultrahigh power density of ~1191 W cm-3. It is noteworthy that these values obtained are among the best values for carbon-based MSCs reported to date.
Article
Full-text available
Significance Microscale batteries can deliver energy at the actual point of energy usage, providing capabilities for miniaturizing electronic devices and enhancing their performance. Here, we demonstrate a high-performance microbattery suitable for large-scale on-chip integration with both microelectromechanical and complementary metal-oxide–semiconductor (CMOS) devices. Enabled by a 3D holographic patterning technique, the battery possesses well-defined, periodically mesostructured porous electrodes. Such battery architectures offer both high energy and high power, and the 3D holographic patterning technique offers exceptional control of the electrode’s structural parameters, enabling customized energy and power for specific applications.
Article
A 3D printing approach is first developed to fabricate quasi‐solid‐state asymmetric micro‐supercapacitors to simultaneously realize the efficient patterning and ultrahigh areal energy density. Typically, cathode, anode, and electrolyte inks with high viscosities and shear‐thinning rheological behaviors are first prepared and 3D printed individually on the substrates. The 3D printed asymmetric micro‐supercapacitor with interdigitated electrodes exhibits excellent structural integrity, a large areal mass loading of 3.1 mg cm−2, and a wide electrochemical potential window of 1.6 V. Consequently, this 3D printed asymmetric micro‐supercapacitor displays an ultrahigh areal capacitance of 207.9 mF cm−2. More importantly, an areal energy density of 73.9 µWh cm−2 is obtained, superior to most reported interdigitated micro‐supercapacitors. It is believed that the efficient 3D printing strategy can be used to construct various asymmetric micro‐supercapacitors to promote the integration in on‐chip energy storage systems. A three‐dimensional (3D) printing approach was developed to fabricate quasi‐solid‐state asymmetric microsupercapacitors. With a high mass loading of 3.1 mg cm−2 and a wide voltage window of 1.6 V, the microsupercapacitor exhibits an ultrahigh areal energy density of 73.9 µWh cm−2. Such an efficient strategy to construct asymmetric microsupercapacitors will promote the integration in on‐chip energy storage systems.
Article
Rational engineering and simplified fabrication of high-energy micro-supercapacitors (MSCs) using graphene and other 2D nanosheets are of great value for flexible and integrated electronics. In this work, we develop one-step mask-assisted simplified fabrication of high-energy MSCs based on the interdigital hybrid electrode (PG) patterns of stacking high quality phosphorene nanosheets and electrochemically exfoliated graphene in ionic liquid electrolyte (denoted as PG-MSCs). The hybrid PG films with interdigital patterns were directly manufactured by layer-by-layer deposition of phosphorene and graphene nanosheets with the assistance of a customized interdigital mask, and directly transferred onto a flexible substrate. The resultant patterned PG films present outstanding uniformity, flexibility, conductivity (319 S cm-1) and structural integration, which can directly serve as binder- and additive-free flexible electrodes for MSCs. Remarkably, PG-MSCs deliver remarkable energy density of 11.6 mWh cm-3, which outperforms the most nanocarbon-based MSCs and thin-film batteries. Moreover, our PG-MSCs show outstanding flexibility and stable performance with slightly capacitance fluctuation even under highly folded states. In addition, our simplified mask-assisted strategy for PG-MSCs is highly flexible for simplified production of parallel and serial interconnected modular power sources, without need of conventional metal-based interconnects and contacts, for designable integrated circuits with high output current and voltage.
Article
Solid-state in-plane microsupercapacitor is realized on the woven textile substrate for the first time with mask-aided electroless coating of Ni and hydrothermal reduction of graphene oxides. High capacitive performances, excellent flexibility, and versatile capability in fashionable/comfortable designs are obtained for the application in smart textiles or wearable electronics.
Article
Highly porous mullite-based thermal insulators were fabricated using a gelation freezing technique. In this process, a honeycomb-like arrangement of pores is created by forming ice crystals within a gel containing dispersed mullite particles, followed by sublimation of the ice under vacuum and subsequent sintering. The uniform arrangement of pore channels and cell walls results in a mullite insulator with a very high compressive strength, even for porosities of up to 91.5 vol%, together with low thermal conductivity. The mechanical strength of this insulator was found to be strongly influenced by the amount of antifreeze protein, the sintering additive in the raw mixture, and the sintering temperature. The use of 0.25 wt% antifreeze protein with the additive 8Y–ZrO2 gave a compressive strength of 11.3 MPa, a porosity of 89.1 vol%, and a thermal conductivity of 0.28 W/m K. This approach can be used to produce macrocellular insulators with specific porosity, thermal conductivity and mechanical properties, suitable for a variety of industrial applications.
Article
Flexible power for flexible electronics A challenge for flexible electronics is to couple devices with power sources that are also flexible. Ideally, they could also be processed in a way that is compatible with current microfabrication technologies. Huang et al. deposited a relatively thick layer of TiC on top of an oxide-coated Si film. After chlorination, most, but importantly not all, of the TiC was converted into a porous carbon film that could be turned into an electrochemical capacitor. The carbon films were highly flexible, and the residual TiC acted as a stress buffer with the underlying Si film. The films could be separated from the Si to form free-floating films, with the TiC providing a support layer. Science , this issue p. 691
Article
Significance Batteries run just about everything portable in our lives such as smartphones, tablets, computers, etc. Although we have become accustomed to the rapid improvement of portable electronics, the slow development of batteries is holding back technological progress. Thus, it is imperative to develop new energy storage devices that are compact, reliable, and energy dense, charge quickly, and possess both long cycle life and calendar life. Here, we developed hybrid supercapacitors that can store as much charge as a lead acid battery, yet they can be recharged in seconds compared with hours for conventional batteries.