You are on page 1of 9

www.particle-journal.

com

www.MaterialsViews.com

FULL PAPER

Fabrication of Cu@MxOy (M = Cu, Mn, Co, Fe) Nanocable


Arrays for Lithium-Ion Batteries with Long Cycle Lives and
High Rate Capabilities
Tianou He, Xingtai Qin, Guang Yang, and Mingshang Jin*
structural particles) has been routinely
adopted to accommodate the local volume
changes, as well as shorten the local diffusion path for ions and electrons, thereby
improving the batterys cycling ability.[710]
Actually, compared to the structural issues
of these types of current collectors, the
problems associated with conventional
planar current collectors are, to some
extent, even more troublesome and also
deserve attention. If a planar current collector is utilized, a rather thick layer of
active material has to be spread on top to
obtain a high energy density. As a result,
kinetic limitations become more acute
due to the macroscopically long diffusion
path. Moreover, because of the disconnection between the current collector and the
upper layers of particles, these particles
will ultimately begin to peel off despite
their delicately designed structures. Consequently, it is difficult
for this class of electrodes to retain a long cycle life. These drawbacks thus prompted the employment of three dimensional (3D)
current collectors, which favor a reduction in coating thickness
while retaining a comparable energy density. More importantly,
the shortened diffusion path allows a fast lithium-ion flux across
the interface and, at the same time, the sufficient interspace can
effectively buffer the volume change.[11]
The key to the development of efficient 3D current collectors is the integrity of the electrodes, which is principally determined by the mechanical strength of the 3D matrix. Moreover,
the intrinsic conductivity of a 3D current collector is crucial for
the rate performance associated with lithium transportation.
From these two points of view, metallic 3D current collectors are
quite advantageous owing to their inherently excellent electric
conductivity and great mechanical strength that favors a uniform
coating and firm attachment of the active materials. To date, a
number of metallic current collectors has been developed and
introduced to enhance the performance of LIBs. For example,
Haag et al. deposited an ultrathin layer of amorphous SnO2 on a
commercial free-standing Ni foam 3D current collector by atomic
layer deposition (ALD) and achieved a high capacity retention for
100 cycles owing to the firm attachment of the particles on the Ni
foam surface.[12] It is worth mentioning that the ALD technique
has also been employed for the uniform coating of active materials on other 3D current collectors such as Ni nanowires (NWs)
and aluminum nanorods.[13,14] Other materials, such as copper

A new strategy is reported to fabricate Cu@MxOy (M = Cu, Mn, Co, Fe)


nanocable arrays using five-fold twinned copper (Cu) nanowire (NW) arrays as
starting materials, to promote both the cycling stability and high rate capability
of MxOy as anodes for LIBs. Conductive Cu NW arrays were synthesized on
Cu foil via chemical vapor deposition (CVD), followed by the oxidation of their
surface so as to form Cu@Cu2O nanocable arrays. The thickness of the active
material (Cu2O) on the Cu NW arrays can be tuned from 20 nm to 160 nm
by simply controlling the oxidation time. Based on this accurate control, the
optimal coating thickness of Cu2O was determined to be around 35 nm.
Additionally, the Cu2O active material shell can be easily transformed to other
metal oxides with even higher specific capacities via a coordinating etching
strategy based on Pearsons principle, resulting in Cu@MxOy nanocable arrays
(M = Mn, Co, Fe). When applied as electrodes for LIBs, these 3D electrodes
show long cycle lives (over 300 cycles) and high rate capabilities.

1. Introduction
Due to the low cost, low toxicity, widespread availability, and high
theoretical capacity, metal oxides are the most promising materials for anodes in lithium-ion batteries (LIBs).[14] Nevertheless,
the poor lithium-transport kinetics and short cycle life have largely
hindered the practical implementation of metal oxides in LIBs.
These problems are generally considered to arise from the long
diffusion path and particle pulverization induced by the enormous
stress generated from the large volume variations that accompany
the repeated lithium insertion/extraction processes.[5,6] Incorporating particles with local void spaces (e.g., hollow and/or porous

T. He, X. Qin, Prof. M. Jin


Frontier Institute of Science and Technology
and Center for Advancing Materials
Performance from the Nanoscale (CAMP-Nano)
Xian Jiaotong University
Xian, Shaanxi 710054, P. R. China
E-mail: jinm@mail.xjtu.edu.cn
Prof. G. Yang
Electronic Materials Research Laboratory
Key Laboratory of the Ministry of Education
& International Center for Dielectric Research
Xian Jiaotong University
Xian 710049, P. R. China

DOI: 10.1002/ppsc.201500125

Part. Part. Syst. Charact. 2015, 32, 10831091

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

1083

FULL PAPER

www.particle-journal.com

www.MaterialsViews.com

pillar arrays and gold nanoporous nanorods, have also been


reported as 3D current collectors that lead to enhanced rate capabilities of LIBs.[15] Although these 3D current collectors have successfully been fabricated, the coating of the active materials was
still limited to the ALD or sputtering technique, which are more
likely to induce a non-uniform deposition of the active materials
on the sub-units of the 3D current collectors. It remains a challenge to uniformly coat the active materials on the 3D current
collector. However, optimizing the uniform coating of active
materials could lead to unprecedented improvements in the
Li-ion diffusion kinetics in the entire device. But, despite earlier
achievements, the optimal coating amount of the active materials
remains unsolved for 3D electrodes, which is crucial for both
power storage and lithium exchange. A thin layer of active materials enables the faster movement of ions and electrons but, at
the same time, it adds a smaller ability for power storage due to
the low packing density of the active materials compared to their
solid-state counterparts. Therefore, an ideal depositing thickness
should simultaneously meet the requirements of optimal lithium
kinetics and high energy density, which, so far, has not been elaborately studied due to the difficulties in achieving uniform coatings of the active material. To attain insight into the dependency
of the cycling performance of these 3D electrodes on a certain
amount of active material, a convenient tuning of the coating
thickness must be achieved. Collectively, it is imperative to simplify the production of metallic 3D current collectors, as well as
the subsequent coating of the active material (metal oxides in
this case) with a tunable thickness, which is of great significance
for both academic investigations and industrial considerations.
Herein, we report on the use of five-fold twinned copper
(Cu) nanowire (NW) arrays as starting materials to fabricate
Cu@MxOy (M = Cu, Mn, Co, Fe) nanocable arrays to promote
both the cycling stability and high rate capability of MxOy as an

anode for LIBs. The conductive Cu NW arrays were synthesized


on Cu foil via a chemical vapor deposition (CVD) technique, followed by the oxidation of their surface so as to form Cu@Cu2O
nanocable arrays. The five-fold twinned Cu core and the strong
binding of Cu2O, as well as the space between the nanocables
can provide structural reinforcement to overcome the mechanical rupture during volume changes of active materials and
enhance the effective electron conduction. Furthermore, the
thickness of the active material (Cu2O) on the Cu NW arrays
can be tuned by simply controlling the oxidation time. Based
on this accurate control, we were able to determine the optimal
coating thickness of Cu2O. Additionally, the Cu2O active material shell could be easily transformed to include other metal
oxides with even higher specific capacities via a coordinating
etching strategy based on Pearsons principle, resulting in
Cu@MxOy nanocable arrays (M = Mn, Co, Fe). When applied
as electrodes for LIBs, these 3D electrodes show long cycle lives
(over 300 cycles) and high rate capabilities.

2. Results and Discussion


2.1. Cu Nanowire Arrays
In this study, instead of using the anodic aluminum oxide (AAO)
template-assisted technique, we employed a chemical vapor deposition (CVD) technique with a modified regular tube furnace as
the synthesis equipment (Figure S1, Supporting Information)
to fabricate Cu NW arrays. Figure 1 shows field-emission scanning electron microscopy (FESEM) images and an X-ray diffraction (XRD) pattern of Cu NW arrays grown on Cu substrates. As
can be seen, the NWs are uniform both in diameter and length.
Figure 1c clearly shows that the Cu NWs have a five-fold twinned

Figure 1. a-c) FESEM images of Cu NW arrays, and d) XRD pattern of Cu NW arrays on Cu foil.

1084

wileyonlinelibrary.com

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Part. Part. Syst. Charact. 2015, 32, 10831091

www.particle-journal.com

www.MaterialsViews.com

2.2. Cu@Cu2O Nanocable Arrays


Due to the well-defined five-fold twinned structure of the Cu
NWs, it was possible to modify and functionalize these NWs
with different properties, for a range of applications. For LIBs,
for example, by simply oxidizing the as-prepared Cu NWs at
200 C in air for 10 minutes, Cu2O could be prepared on the
surface of the Cu NWs, resulting in Cu@Cu2O nanocable
arrays (Figure 2ac). As can be seen in Figure 2d, the surface
of the preformed Cu NWs is relatively smooth. After oxidation
for 10 minutes in air (Figure 2e), the surface of the Cu NWs
was more rough due to the formation of small Cu2O nanoparticles, which was further confirmed by XRD characterization
(the blue curve in Figure 2g). EDS elemental mapping of the
nanowires also revealed the nanocable structure of Cu@Cu2O
(Figure S4, Supporting Information). It can be seen from EDS
that Cu is widely distributed in the whole area, as the Cu NWs
are grown on a Cu substrate. However, the wide distribution
of O clearly indicates the Cu@Cu2O coreshell structure. In
order to confirm the well-designed co-axial structure of the
Cu@Cu2O nanocables, the Cu2O shells were removed by
1 M Na2S2O3 solution. As shown in Figure 2f,g, the NW cores
survived the etching with Na2S2O3 solution and, more importantly, retained their five-fold twinned configuration, implying
that the Cu NW cores attained a great mechanical strength
(actually, this also applied to the 3D current collectors after
oxidation).

FULL PAPER

structure. The XRD pattern (Figure 1d) indicates that the relative
intensity of the Cu (220) diffraction is much stronger than that
of Cu (200), suggesting that the vertical orientation of the NWs
is parallel to the [110] axial, which is in agreement with previous
reports on the growth direction of five-fold twinned Cu NWs.[16,17]
To our knowledge, the five-fold twinned structure of metal NWs
benefits the mechanical strength of NWs,[18,19] which favors the
subsequent uniform coating of active materials on their surfaces.
In comparison, Cu NWs prepared in AAO templates tend to
exhibit a structure that is a mixture of amorphous and crystalline
domains because of the existence of impurities resulting from
AAO templates and their fabrication process (e.g., Al and P),
making it hard to remain intact during the oxidation and retain
its mechanical strength.[20] Therefore, a much better mechanical
strength could be expected from Cu NWs synthesized via the
CVD technique in this work. Noteworthy and interestingly, the
five-fold twinned Cu NWs were also successfully obtained on
foils other than Cu, such as glass sheets, a silicon wafer, and even
an adhesive tape (Figure S2, Supporting Information), verifying
the versatility of this method. Ex situ SEM studies at the early
stages of the growth indicated that a layer of small Cu particles
with well-recognized five-fold twinned structure formed prior
to the formation of the NWs (Figure S3, Supporting Information). These particles then served as seeds for further growth into
NWs. Governed by this growth manner, these five-fold twinned
Cu NWs can be fabricated on various substrates, since these Cu
seeds can be easily deposited on different substrates.

Figure 2. ac) Illustrations and df) SEM images of Cu NW arrays before oxidation, after oxidation, and after removal of Cu2O shells. g) XRD patterns
of Cu NWs and Cu@Cu2O nanocables. The scale bars represent 200 nm.

Part. Part. Syst. Charact. 2015, 32, 10831091

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

1085

www.particle-journal.com

FULL PAPER

www.MaterialsViews.com

Figure 3. Galvanostatic chargedischarge profiles between 0.05 and 3 V vs Li/Li+ for the 1st, 2nd, 10th, and 80th cycles at a rate of 0.5 C of a) Cu@Cu2O
nanocable array electrode and b) planar Cu/Cu2O electrode. Discharge capacities and Coulombic efficiencies of c) Cu@Cu2O nanocable array electrode and
d) planar Cu/Cu2O electrode.

This mechanically stable configuration undoubtedly has


great potential to be used as the electrode for LIBs, significantly
increasing the cycle times of batteries. Therefore, we evaluated the Cu@Cu2O nanocable arrays as anodes for LIBs. We
benchmarked the capability of our Cu@Cu2O nanocable array
electrode against a planar Cu/Cu2O electrode (Figure S5, Supporting Information). Figure 3a,b displays the first, second,
10th, and 80th dischargecharge curves of a Cu@Cu2O nanocable array electrode oxidized at 200 C for 10 min and of a
planar Cu/Cu2O electrode at a current density of 200 mA g1
(ca. 0.5 C). The first discharge curve of the Cu@Cu2O nanocable array electrode has an extended potential plateau at
around 1.5 V, followed by a sloping potential at around
0.5 V, which was mainly caused by the electrochemical reaction
pathway. The cyclic voltammetry (CV) curve of the Cu@Cu2O
nanocable array electrode clearly exhibited two peaks at around
0.5 V and around 1.5 V (Figure S6, Supporting Information).
These two peaks could be ascribed to the formation of a solid
electrolyte interface (SEI) and the reduction of Cu2O to Cu,
respectively.[21] It should be pointed out that the formation
of a SEI is largely related to the wide voltage window used
(0.053 V), which is generally considered to cause a dramatically reduced cycling performance due to the harsh conditions
for lithium extraction.[22] Figure 3c,d presents the cycling performance of the 3D Cu@Cu2O nanocable array electrode and a
Cu/Cu2O planar electrode. The initial Coulombic efficiency of
the 3D electrode was 67.3%, which increased to over 98% in the
following 80 cycles. In contrast, the initial Coulombic efficiency
of the planar electrode was 51.7%. The capacity loss for the 3D
Cu@Cu2O nanocable array electrode is clearly negligible over
80 cycles whereas the capacity of the planar Cu/Cu2O electrode
decayed to only 42% of its initial value after 80 cycles.

1086

wileyonlinelibrary.com

To figure out the cause for this difference, we conducted


ex situ studies of both electrodes after 80 cycles using SEM
characterization. As shown in Figure S7a (Supporting Information), a disintegration of the planar Cu/Cu2O electrode was
seen after cycling for 80 times and the Cu2O nanoparticles were
found to have broken up into nanoparticles of even smaller
sizes, which is similar to previous reports.[21] Accordingly, the
irreversible capacity loss of planar Cu/Cu2O electrodes can
mainly be attributed to the formation of a SEI layer and to
partially irreversible electrochemical reactions during further
cycling because of the disintegration of Cu2O nanoparticles
and the following severe aggregation. On the contrary, owing
to the ultrahigh mechanical strength of the Cu NW cores,[23,24]
the micrometer-sized nanocable arrays were able to minimize
bulk distortion during repeated charging/discharging procedures, which ensured the electrodes integrity (Figure S7b).
Moreover, as shown in Figure S8 in the Supporting Information, electrochemical impedance spectroscopy (EIS) indicated
that the impedance decreased after cycling, verifying that the
nanostructured Cu@Cu2O nanocable array electrode improved
the electronic/ionic conductivity of the Cu2O/Cu0/Li2O matrix.
This can be interpreted as the easy wetting and activation of the
surface materials because of the adequate electrolyte/Cu2O and
Cu2O/Cu contact caused by the large surface area of the nanowires array and the thin layer of Cu2O.[25,26] These advantages are
responsible for the superior cycling ability of the 3D electrode.
Moreover, an excellent rate capability was also achieved for the
durable 3D material. As shown in Figure 4a, when the current
rate was increased from 0.5 C to 8 C, the 3D electrode could still
recover over 60% of the total capacity. Significantly, when the
current rate was reduced back to 0.5 C, the total capacity was
fully restored. On the contrary, only 30% of the total capacity

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Part. Part. Syst. Charact. 2015, 32, 10831091

www.particle-journal.com

www.MaterialsViews.com

FULL PAPER

Figure 4. Cycling performance of (a) Cu@Cu2O nanocable array electrode


and (b) Cu/Cu2O planar electrode at various rates.

could be achieved for the Cu/Cu2O planar electrode when the


current rate was increased to 8 C (Figure 4b).
The firm attachment of a Cu2O shell on the Cu NWs is thus
crucial for the excellent cycling performance. Considering the
fact that the Cu2O shell of the above-mentioned 3D electrode
was formed by oxidation, it is rational to expect that a tunable
thickness of the shell can be realized. Ex-situ SEM studies of
Cu@Cu2O nanocables that had been oxidized for different
amounts of time from Cu NWs enabled us to investigate the
evolution from Cu NWs to Cu@Cu2O nanocables. Figure S8a
in the Supporting Information clearly presents the formation
of slight protuberances at the outline of the five-fold twinned
Cu NWs after oxidation for 5 min at 200 C. The protuberances grew larger if the oxidation time was prolonged to 30 min
(Figure S9b, Supporting Information). After removal of the
Cu2O shell, as shown in Figure S9c, a distinct five-petals structure could be observed. It therefore can be concluded that the
NWs are preferably oxidized from the twinned planes, which
were indexed to be Cu (111) facets, as depicted in Figure S9d.
This is similar to the reports by Xia et al.,[27] who proposed that
oxygen preferably etched metal nanocrystals from their crystal
boundaries in the solution phase (e.g., the twinned planes).
Theoretically, twinned planes are generally more active than
low-index crystal planes. As a result, etching would preferably
start from the twinned planes, which is also valid for the oxidation process of the Cu NWs in our experiments.
This manner of oxidation for the Cu NWs also suggests that
the thickness of the Cu2O shells coated on the Cu NW arrays
could be easily tuned by varying the oxidation time. Figure 5ae
shows SEM images of a series of nanocable arrays obtained at different oxidation times. Evidently, the diameters of the nanocables
increase with the reaction time, implying a continued increase in
the thickness of the Cu2O shells coated on the Cu NWs. The XRD

Figure 5. ae) Morphologies of Cu NWs oxidized for 1 min, 5 min, 10 min, 30 min, and 60 min. f) XRD patterns of Cu NWs oxidized for different times.

Part. Part. Syst. Charact. 2015, 32, 10831091

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

1087

FULL PAPER

www.particle-journal.com

1088

www.MaterialsViews.com

patterns show that the intensity of the {111} diffraction of the


Cu2O shells experienced a continuous enhancement whereas
that of the {220} diffraction of Cu shows a continual decrease.
The change in the relative intensity of {111}/{220} also confirms
that the ratio of Cu2O to Cu is increasing. In addition, the SEM
images indicate that the Cu2O shell maintains its uniform distribution along the NW surfaces regardless of the oxidation time,
confirming the reliability and accessibility of our method for the
uniform coating of active materials on 3D current collectors. For
practical considerations, we determined the thicknesses of the
Cu2O shells on all electrodes by taking the difference between
the diameter of the Cu@Cu2O nanocables and that of the core
after the Cu2O shell was removed by Na2S2O3. As demonstrated
in Figure S10 (Supporting Information), the thickness of the
active material could be controlled continuously ranging from
20 nm to 35, 60, and 160 nm, corresponding to oxidation times
of 5, 10, 30, and 60 min, respectively.
Thanks to the convenient tuning of the Cu2O shell thickness, we were able to investigate the dependency of the electrochemical performance of batteries on the thickness of the
active materials deposited on 3D current collectors. As shown
in Figure 6a, when normalized to the cross-sectional area of the
current collector, the sample oxidized for 10 min, corresponding
to a shell thickness of 35 nm, attained the highest specific
capacity. Therefore, an adequately thin film of active material
would promote outstanding current collector/active material
surface contacts, creating an non-impeded path for ions and
electrons and thus an excellent high rate cycling performance.
From this perspective, increasing the thickness of the active
layer would benefit the cycling performance as the energy density is improved. However, with increasing thickness new penalties related to kinetic limitations will appear bearing in mind
that the thicker the active layer, the longer and more rugged the
path for ions and electrons becomes. Moreover, when a large
amount of active material is spread on the current collector, a
fraction of the particles is actually less active or even inactive
because of the lack of contact between these particles and either
the current collector or the electrolyte. Most anode materials to
date are composed of semiconductors with poor electrical conductivity. Thus, increasing the amount of active materials will
undoubtedly affect the electron/ion conductivity by creating too
long a diffusion path. Here, a Cu2O shell thickness of 35 nm
simultaneously meets the requirements of energy storage and
fluent electron/ion transportation. We believe that this result
can also be extended to other 3D current collector-assisted metal
oxide electrodes, given the fact that most metal oxides undergo
similar pathways for chemical reactions and lithium-ion insertion/extraction when utilized as electrodes for LIBs.
We also tested the cycle lives of our as-prepared electrodes.
Amazingly, as shown in Figure 6b, regardless of the specific
capacity, all batteries manifested a considerably stable capacity
retention for over 300 cycles, which is much better than the cycle
life of other 3D current collector-assisted electrodes and Cu2Orelated electrodes that have been reported before.[21,28] These high
stabilities could be ascribed to the ultra-stable co-axial structure of
the Cu@Cu2O nanocable arrays. We further conducted post-utilization investigations of all electrodes after 300 charge/discharge
cycles. As displayed in Figure S11 (Supporting Information), the
initial Cu2O shells on Cu NWs turned into isolated nanoparticles,

wileyonlinelibrary.com

Figure 6. a) Specific capacities and b) capacity retentions of Cu@Cu2O


nanocables with different thicknesses of Cu2O. All samples were oxidized
at 200 C, legends of the plots indicate the oxidation times of the samples.

yet still adhered to the Cu NWs after 300 cycles. Moreover, these
reconstructed particles were still distributed densely on the surface of the Cu NWs. It can be seen that instead of the particle
pulverization that happened in many other cases,[2932] a new
layer of active materials composed of Cu2O particles that were
larger than the initial ones had formed and was uniformly coated
on the 3D Cu NWs current collector. In this way, the integrity
of the electrodes was retained to a large extent and the capacities were well maintained. Again, we can attribute this extremely
high stability of our electrodes to the great mechanical strength
of the preformed Cu NWs, which helps to maintain the configuration of the 3D current collector as well as favors a close and
stable interaction between the Cu2O particles and the conductive
NW cores. According to previous reports, the Li2O matrix formed
during repeated Li+ insertion/extraction processes also likely acts
as an adhesive for keeping the particles on the Cu NWs.[13,33]
2.3. Cu@MxOy (M = Mn, Co, Fe) Nanocable Arrays
Although we achieved long cycle lives and high rate capabilities achieved for our 3D electrodes, there were still concerns for
their practical application considering that the specific capacity

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Part. Part. Syst. Charact. 2015, 32, 10831091

www.particle-journal.com

www.MaterialsViews.com

Cu 2 O + xS2 O32 + H2 O [Cu 2 (S2 O32 )x ]2 2x + 2OH

(1)

S2 O32 + H2 O  HS2 O3 + OH

(2)

Mn + + nOH M(OH)n

(3)
+-O2

bond converts into a softsoft


Firstly, the softhard Cu
Cu+-S2O32 bond, forming a soluble phase accompanied by the

production of OH that is also generated by the hydrolysis of


S2O32. Then, Mn+ reacts with OH to form M(OH)n because
of the hard acid nature of Mn+. Finally, M(OH)n dehydrates to
form MxOy after thermal treatment. Figure 7x1 (x = ac) reveals
FESEM images of different Cu@MxOy nanocables (M = Mn,
Co, Fe). The worm-like shell structures of Mn3O4 and Co3O4
agree well with Guos study.[34] It should be pointed out that the
Fe2O3 shell seems thinner than that of the other two oxides.
This is because iron hydroxide is less soluble than manganese
hydroxide and cobalt hydroxide. In other words, the relatively
high pH of the system leads to a faster aggregation of iron
hydroxide and, thus, a thinner layer of Fe2O3. However, we have
shown that the Cu2O shell could successfully be substituted by
other active materials with higher specific capacities.
Figure 7x2 and Figure 7x3 demonstrate the cycling performances and rate capabilities of the 3D Cu@MxOy nanocables
electrodes. A high specific capacity of up to 872.1 mAh g1, which
is approximately 93% of the theoretical capacity, was achieved for
Mn3O4 coated on the Cu NWs. After 300 chargedischarge cycles
at 0.5 C (C = 936 mA g1), a capacity as high as 800 mAh g1 was
retained. This high capacity and long cycle life indicate the fully

FULL PAPER

of Cu2O is relatively low compared to that of other metal oxides.


Inspired by the previous report by Guo et al., which mentions
that metal oxide nanocages can be generated from Cu2O particles based on Pearsons hard and soft acidbase principle,[34] we
designed a feasible approach to the transformation from Cu@
Cu2O nanocables to Cu@MxOy nanocables (M = Mn, Co, Fe)
using the sample that had been oxidized for 10 min as it possessed the highest specific capacity. We selected these metals
for two reasons: i) their oxides are known to have high specific
capacities;[3537] ii) the corresponding metal ions are harder
acid compared to Cu+, which favors the transformation process
depicted in Figure S12 (Supporting Information). The conversion process can be explained as:

Figure 7. SEM images, capacity retentions, and rate capabilities of a) Cu@Mn3O4 nanocables, b) Cu@Co3O4 nanocables, and c) Cu@Fe2O3 nanocables.

Part. Part. Syst. Charact. 2015, 32, 10831091

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

1089

FULL PAPER

www.particle-journal.com

www.MaterialsViews.com

realized and perfectly reversible Li+ insertion/extraction process


reflected in the reaction: 8Li+ + 8e + Mn3O4 3Mn + 4Li2O.[38]
The CV curve also confirms the above discharging process. As
shown in Figure S13a (Supporting Information), the two anodic
peaks at around 1.2 V and 0.6 V can be attributed to the reduction
of MnIII to MnII and the reduction of MnII to Mn0, respectively,[39]
indicating the complete reduction of MnIII Mn0. Similarly,
Co3O4 on Cu NWs also showed a high specific capacity of
801.9 mAh g1, which is slightly lower than the theoretical value
(888 mAh g1). The CV behavior of the Cu@Co3O4 electrode
(Figure S13b) shows two characteristic anodic peaks at around
1.35 V and 0.7 V, corresponding to the reduction of CoIII CoII
and that of CoII Co0, respectively.[40] Therefore, the achieved
capacity can be attributed to the accessible and completely reversible reaction: 8Li+ + 8e + Co3O4 4Li2O + 3Co.[38,41] The capacity
decayed by only 10% after 300 cycles at 0.5 C (C = 888 mA g1),
implying a good cycling stability. The capacity of the Fe2O3 on Cu
NWs was 799.2 mAh g1, which decreased to 689.5 mAh g1 after
300 cycles at 0.5 C (1005 mAh g1). The gap between the achieved
capacity and the theoretical value can be interpreted as follows: as
shown in Figure S13c, two anodic peaks at around 1.6 V and 0.7 V
appear in the CV curve. The former can be assigned to the Li+
insertion into the Fe2O3 structure.[42] The latter can be ascribed to
the reduction of the Fe ion to Fe0.[43] However, the peak position
is at a more negative potential than that in previous reports, indicating an incomplete FeIIIFe0 transformation due to the partial
reoxidation of Fe0 to FeO via 2Li2O + 2Fe 2FeO + 4Li+ + 4e.[44]
Despite these limitations, the relatively high capacity demonstrated
by the Cu@Fe2O3 electrode is still comparable with those reported
previously. Cycling performances at varied current rates revealed
that all these electrodes can recover over 50% of their capacity
when the current rate was increased from 0.5 C to 8 C. After the
current rate was reduced from 32 C back to 0.5 C, over 90% of
the capacities were retained for all electrodes. These high capacities and excellent performances related to cycle lives and rate
capabilities can be explained by two aspects: 1) the intimate interaction between the active materials and the mechanically stable
current collector, which ensures the integrity of the 3D electrodes
and thereby enhances the cycling stabilities; 2) as only a relatively
thin layer of metal oxides is dispersed on the 3D current collector, the outstanding electrical conductivity of Cu NWs makes
them electrochemically active. For instance, Mn3O4 is known to
own an extremely low electrical conductivity,[45] which limits its
achievable capacity because it is less active. In the present work,
the Cu NW arrays, which act as current collectors, can effectively
enhance the mobility of the charge carriers and thus facilitate the
transfer of these charge carriers from the current collector to the
active materials, thus improving the reactivity of the metal oxides
and optimizing the specific capacities.

3. Conclusion
Cu NW arrays were successfully synthesized on different substrates using a CVD technique. Different from conventional
loading methods of active materials, the surface of these Cu
NWs could be oxidized uniformly to form Cu@Cu2O nanocables, which can then be used as electrodes for LIBs. Due to the
well-defined co-axial structure, the electrochemical performance

1090

wileyonlinelibrary.com

of these Cu@Cu2O nanocables exhibited super high cycling


stabilities and rate capabilities. Moreover, the specific capacity
was found to be quite dependent on the thickness of the Cu2O
layer and the optimal coating thickness was determined to be
around 35 nm in our work. Additionally, the Cu2O shell could
be easily replaced by other metal oxides with even higher specific capacities via a coordinating etching strategy based on
Pearsons principle, resulting in Cu@MxOy nanocable arrays
(M = Mn, Co, Fe). When applied as electrodes for LIBs, these
3D electrodes showed long cycle lives (over 300 cycles) and
high rate capabilities. We believe that this work can be further
expanded to other 3D current collectors and work as a reference
for industrial considerations.

4. Experimental Section
Synthesis of Copper Nanowire Arrays: Copper precursor, Cu(etac)
[P(OEt)3]2, was prepared by mixing 600 mg CuCl with 2.0 mL P(OEt)3
in about 3 mL of tetrahydrofuran (THF), followed by adding 0.912 g of
Na(etac) into the suspension, and finally isolation as volatile liquid by
centrifugation. The Cu NWs were grown in a modified hot-wall CVD
system (see Figure S1, Supporting Information). The system consisted
of two heating zones: the anterior for the evaporation of the Cu precursor
and the posterior for the Cu NWs growth where the substrate was placed.
In a typical growth procedure, the precursor evaporation heating zone
was set to 65 C, the growth heating zone was set to 180 C, and the CVD
system was kept under a low pressure of about 0.5 Torr during the whole
growth process using a vacuum pump. The deposition time could be
extended from 2 to 8 h to tune the diameter and length of the Cu NWs.
Fabrication of Cupric Oxide on Copper Nanowires: In a typical synthesis,
Cu@Cu2O nanocable arrays were fabricated by oxidizing the as-prepared
Cu NWs substrate in a preheated muffle furnace under ambient
conditions for 10 min at 200 C. Cu@Cu2O nanocables with various
thicknesses of 20 nm to 160 nm were obtained by varying the oxidation
time ranging from 5 min to 60 min.
Fabrication of Cu@MxOy Nanocable Arrays: Other metal oxides were
prepared employing a coordinating etching strategy based on Guos
report.[24]
Cu@Mn3O4 Nanocables: Cu@Cu2O nanocables and MnCl24H2O
(0.5 mg) were added to 10 mL of an ethanol/water (v/v = 7:3) mixed
solvent in the presence of poly(vinylpyrrolidone) (PVP) (0.3333 g,
Mw = 30 000 g mol1). After the mixture was stirred for 10 min, 4 mL
of Na2S2O3 aqueous solution (0.2 M) was added dropwise. Then the
reaction was carried out at room temperature for 5 min. Cu@Mn(OH)2
nanocables obtained at this stage were then calcined at 400 C for 30 min
under the protection of argon to generate Cu@Mn3O4 nanocables.
Cu@Co3O4 Nanocables: Cu@Cu2O nanocables and CoCl26H2O
(0.8 mg) were added to 10 mL of an ethanol/water (v/v = 1:1) mixed
solvent in the presence of PVP (0.3333 g, Mw = 30 000 g mol1). After
the mixture was stirred for 10 min, 4 mL of Na2S2O3 aqueous solution
(1 M) was added dropwise. Then the reaction was further carried out at
room temperature for 3 min. The Cu@Co(OH)2 nanocables obtained at
this stage were then calcined at 400 C for 30 min under the protection
of argon to generate Cu@Co3O4 nanocables.
Cu@Fe2O3 Nanocables: Cu@Cu2O NWs and FeCl24H2O (1 mg) were
added to 10 mL of deionized water in the presence of PVP (0.3333 g,
Mw = 30 000 g mol1). After the mixture was stirred for 10 min, 1 mL of
Na2S2O3 aqueous solution (1 M) was added dropwise. Then the reaction
was carried out at room temperature for 3 min. The Cu@Fe(OH)3
nanocables obtained at this stage were then calcined at 400 C for 30 min
under the protection of argon to generate Cu@Fe2O3 nanocables.
Preparation of Cu/Cu2O Planar Electrodes: Firstly, Cu particles were
fabricated on a Cu foil following the same procedure as for the synthesis
of the Cu@Cu2O nanocables except that the deposition time was
shortened to 40 min. Then, the Cu particles deposited on Cu foil were

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Part. Part. Syst. Charact. 2015, 32, 10831091

www.particle-journal.com

www.MaterialsViews.com

Supporting Information
Supporting Information is available from the Wiley Online Library or
from the author.

Acknowledgements
T.H. and X.Q. contributed equally to this work. All authors discussed the
results and commented on the paper. The authors declare no competing
financial interest. This work was supported by the National Natural
Science Foundation of China (NSFC, No. 21403160 and 21471123), and
the start-up fund, the Fundamental Research Funds for the Central
Universities provided by Xian Jiaotong University. The authors thank
Prof. G. Yang of Xian Jiaotong University for his help with the SEM
characterization.
Received: July 16, 2015
Revised: September 11, 2015
Published online: November 5, 2015

[1] W. Zhou, C. Cheng, J. Liu, Y. Y. Tay, J. Jiang, X. Jia, J. Zhang, H. Gong,


H. H. Hng, T. Yu, H. J. Fan, Adv. Funct. Mater. 2011, 21, 2439.
[2] L. Wang, D. Wang, Z. Dong, F. Zhang, J. Jin, Small 2014, 10, 998.
[3] Z. Wang, L. Zhou, X. W. Lou, Adv. Mater. 2012, 24, 1903.
[4] Y. Huang, D. Wu, J. Wang, S. Han, L. Lv, F. Zhang, X. Feng, Small
2014, 10, 2226.
[5] H. Zhang, X. Yu, P. V. Braun, Nat. Nanotechnol. 2011, 5, 277.
[6] Y. Yu, L. Gu, C. Zhu, S. Tsukimoto, P. A. van Aken, J. Maier, Adv.
Mater. 2010, 22, 2247.
[7] L. Mai, Q. An, Q. Wei, J. Fei, P. Zhang, X. Xu, Y. Zhao, M. Yan,
W. Wen, L. Xu, Small 2014, 10, 3032.
[8] J. Ye, H. Zhang, R. Yang, X. Li, L. Qi, Small 2010, 6, 296.
[9] K. Brezesinski, J. Haetge, J. Wang, S. Mascotto, C. Reitz, A. Rein,
S. H. Tolbert, J. Perlich, B. Dunn, T. Brezesinski, Small 2011, 7, 407.
[10] L. Zhao, Y. S. Hu, H. Li, Z. Wang, L. Chen, Adv. Mater. 2011, 23, 1385.

Part. Part. Syst. Charact. 2015, 32, 10831091

[11] S. J. Ding, Z. Y. Wang, S. Madhavib, X. W. Lou, J. Mater. Chem. 2011,


21, 13860.
[12] J. M. Haag, G. Pattanaik, M. F. Durstock, Adv. Mater. 2013, 25, 3238.
[13] W. Wang, M. Tian, A. Abdulagatov, S. M. George, Y. C. Lee, R. Yang,
Nano Lett. 2012, 12, 655.
[14] S. K. Cheah, E. Perre, M. Rooth, M. Fondell, A. Harsta, L. Nyholm,
M. Boman, T. Gustafsson, J. Lu, P. Simon, K. Edstrom, Nano Lett.
2009, 9, 3230.
[15] P. L. Taberna, S. Mitra, P. Poizot, P. Simon, J. M. Tarascon, Nat.
Mater. 2006, 5, 567.
[16] C. Kim, W. Gu, M. Briceno, I. M. Robertson, H. Choi, K. Kim, Adv.
Mater. 2008, 20, 1859.
[17] H. Choi, S. H. Park, J. Am. Chem. Soc. 2004, 126, 6248.
[18] A. M. Leach, M. McDowell, K. Gall, Adv. Funct. Mater. 2007, 17, 43.
[19] B. Wu, A. Heidelberg, J. J. Boland, J. E. Sader, X. M. Sun, Y. D. Li,
Nano Lett. 2006, 6, 468.
[20] M. I. Irshad, F. Ahmad, N. M. Mohamed, M. Z. Abdullah, Int.
J. Electrochem. Sci. 2014, 9, 2548.
[21] C. Q. Zhang, J. P. Tu, X. H. Huang, Y. F. Yuan, X. T. Chen, F. Mao,
J. Alloys Compounds 2007, 441, 52.
[22] X. Wen. Lou, C. M. Li, L. A. Archer, Adv. Mater. 2009, 21, 2536.
[23] E. W. Wong, P. E. Sheehan, C. M. Lieber, Science 1997, 277, 1971.
[24] B. Wu, A. Heidelberg, J. J. Boland, Nat. Mater. 2005, 4, 525.
[25] Y. Oumellal, N. Delpuech, D. Mazouzi, N. Dupr, J. Gaubicher,
P. Moreau, P. Soudan, B. Lestriez, D. Guyomard, J. Mater. Chem.
2011, 21, 6201.
[26] L. Wang, Y. Yu, P. C. Chen, D. W. Zhang, C. H. Chen, J. Power
Sources 2008, 183, 717.
[27] Y. Xia, Y. Xiong, B. Lim, S. E. Skrabalak, Angew. Chem. 2009, 121, 62;
Angew. Chem. Int. Ed. 2009, 48, 60.
[28] S. Grugeon, S. Laruelle, R. Herrera-Urbina, L. Dupont, P. Poizot,
J.-M. Tarascona, J. Electrochem. Soc. 2001, 148, A285.
[29] Y. Liu, N. S. Hudak, D. L. Huber, S. J. Limmer, J. P. Sullivan,
J. Y. Huang, Nano Lett. 2011, 11, 4188.
[30] G. M. Zhou, D. W. Wang, F. Li, L. L. Zhang, N. Li, Z. S. Wu, L. Wen,
G. Q. Lu, H. M. Cheng, Chem. Mater. 2010, 22, 5306.
[31] D. H. Nam, K. S. Hong, S. J. Lim, H. S. Kwon, J. Power Sources 2014,
247, 423.
[32] P. C. Lian, X. F. Zhu, S. Z. Liang, Z. Lia, W. S. Yang, H. H. Wang,
Electrochim. Acta 2011, 56, 4532.
[33] C. M. Wang, W. Xu, J. Liu, J. G. Zhang, L. V. Saraf, B. W. Arey,
D. Choi, Z. G. Yang, J. Xiao, S. Thevuthasan, D. R. Baer, Nano Lett.
2011, 11, 1874.
[34] J. W. Nai, Y. Tian, X. Guan, L. Guo, J. Am. Chem. Soc. 2013, 135, 16082.
[35] J. Gao, M. A. Lowe, H. D. Abruna, Chem. Mater. 2011, 23, 3223.
[36] R. Wu, X. Qian, X. Rui, H. Liu, B. Yadian, K. Zhou, J. Wei, Q. Yan,
X. Feng, Y. Long, L. Wang, Y. Huang, Small 2014, 10, 1932.
[37] Y. Chen, B. Song, X. Tang, L. Lu, J. Xue, Small 2014, 10, 1536.
[38] Z. Wu, W. Ren, L. Wen, L. Gao, J. Zhao, Z. Chen, G. Zhou, F. Li,
H. Cheng, ACS Nano 2010, 4, 3187.
[39] D. Pasero, N. Reeves, A. R. West, J. Power Sources 2005, 141, 156.
[40] W. Y. Li, L. N. Xu, J. Chen, Adv. Funct. Mater. 2005, 15, 851.
[41] J. Jiang, J. P. Liu, R. M. Ding, X. X. Ji, Y. Y. Hu, X. Li, A. Z. Hu, F. Wu,
Z. H. Zhu, X. T. Huang, J. Phys. Chem. C 2010, 114, 929.
[42] J. Morales, L. Sanchez, F. Martin, F. Berry, X. L. Ren, J. Electrochem.
Soc. 2005, 152, A1748.
[43] X. J. Zhu, Y. W. Zhu, S. Murali, M. D. Stoller, R. S. Ruoff, ACS Nano
2011, 4, 3333.
[44] M. V. Reddy, T. Yu, C. H. Sow, Z. X. Shen, C. T. Lim, G. V. Subba
Rao, B. V. R. Chowdari, Adv. Funct. Mater. 2007, 17, 2792.
[45] H. Wang, L. F. Cui, Y. Yang, H. S. Casalongue, J. T. Robinson,
Y. Liang, Y. Cui, H. Dai, J. Am. Chem. Soc. 2010, 132, 13978.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

FULL PAPER

oxidized in the preheated muffle furnace under ambient conditions for


10 min at 200 C.
Determination of the Weight of the Active Materials: The as-prepared
Cu@Cu2O nanocables array grown on Cu foil was cut into several pieces
with the same area. One of the pieces was used to determine the mass
of the Cu cores array after removing the Cu2O shell with Na2S2O3. The
weight of the active materials on the 3D current collector was estimated
by calculating the weight difference between the Cu@MxOy nanocable
arrays and the Cu core arrays.
Electrochemical Measurements: The as-prepared Cu@Cu2O nanocables
on copper foil (size 1 cm 1 cm) were used as the cathode. CR 2032
coin-type cells were assembled in an Ar-filled glove box, using lithium
metal foil as the counter electrode. The electrolyte was 1 M LiPF6 in
a mixture of ethylene carbonate (EC) and diethyl carbonate (DEC)
(v/v = 1:1). All cells were galvanostatically discharged and charged in
the range of 0.013.0 V at various current densities. Cyclic voltammetry
measurements of the electrodes were performed on a CHI660B
electrochemical workstation with a scan rate of 0.5 mV s1 between
0.01 and 3.0 V (versus Li+/Li).
Characterizations: Powder X-ray diffraction (XRD) was performed
using a diffractometer (DMAX/A, Rigaku) operated at 40 kV and 30 mA
with Cu K radiation. The morphologies of all samples were characterized
with a Nova Nano-SEM 230 field-emission microscope (FEI, Hillsboro,
OR) operated at 30 kV.

1091

You might also like