Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Power Sources 458 (2020) 228005

Contents lists available at ScienceDirect

Journal of Power Sources


journal homepage: www.elsevier.com/locate/jpowsour

Oxygen vacancies enhance supercapacitive performance of CuCo2O4 in


high-energy-density asymmetric supercapacitors
Yamin Feng a, b, Weifeng Liu c, Yin Wang d, Wenning Gao a, Jitao Li a, Kuili Liu a, **,
Xiaoping Wang a, Jian Jiang b, *
a
College of Physics and Telecommunication Engineering, Zhoukou Normal University, 466001, Zhoukou, PR China
b
Chongqing Key Laboratory for Advanced Materials & Technologies of Clean Energies, Chongqing, 400715, PR China
c
School of Physics, Huazhong University of Science and Technology (HUST), Wuhan, 430074, PR China
d
Hubei Key Laboratory of Low Dimensional Optoelectronic Materials and Devices, Hubei University of Arts and Science, Xiangyang, 441053, PR China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

� Mesoporous CuCo2O4 nanoflowers with


enriched oxygen vacancies are
synthesised.
� The sample exhibits abundant active
sites and enhanced electronic
conductivity.
� The electrode delivers high specific
capacitance and excellent cyclic
performance.
� The device displays a high energy den­
sity of 58.7 Wh kg 1 (at 800 W kg 1).

A R T I C L E I N F O A B S T R A C T

Keywords: Although the emerging CuCo2O4 electrode material proves to be promising for use in energy-storage applica­
Oxygen vacancies tions, its slow reaction kinetics and poor conductivity limit its broad utilisation. To address this challenging
CuCo2O4 fundamental issue, herein, oxygen-vacancy-enriched CuCo2O4 nanoflowers is prepared using a facile hydro­
Electrochemical performances
thermal method followed by thermal treatment in a hypoxic atmosphere. Because of the presence of oxygen
Supercapacitor
defect sites and impurity bands, such flower-like CuCo2O4 nanomaterials with large specific surface areas
demonstrate much superior electrical conductivity and favourable hydrophilic properties, which are highly
encouraging for supercapacitor applications. Impressively, when evaluated as an active electrode material, it
exhibits a remarkable specific capacitance (1006 F g 1 at 1 A g 1, i.e., 1.2 F cm 2 at 1.2 mA cm 2), excellent rate
capability (69.4% capacitance retention at 20 A g 1) and ultra-long cycling lifespan (85.5% specific capacitance
retention after 10,000 cycles). Moreover, when being paired with activated carbon, the quasi-solid-state asym­
metric supercapacitors provide a maximum energy density of 58.7 Wh kg 1 at a power density of 800 W kg 1
and extraordinary cycling stability (71.2% retention after 10,000 cycles). These results firmly verify that the
proper incorporation of oxygen vacancies into metal oxides provides a new efficient pathway to advance elec­
trode behaviours.

* Corresponding author.,
** Corresponding author.
E-mail addresses: liukuili@163.com (K. Liu), jjiang@swu.edu.cn (J. Jiang).

https://doi.org/10.1016/j.jpowsour.2020.228005
Received 16 November 2019; Received in revised form 29 January 2020; Accepted 8 March 2020
Available online 20 March 2020
0378-7753/© 2020 Elsevier B.V. All rights reserved.
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

1. Introduction cm 2), demonstrating high rate capability and cycling stability (greater
than 85% capacitance retention after 10,000 cycles even at 50 mV s 1).
With an increasing demand for electric vehicles and portable elec­ Furthermore, an asymmetric supercapacitor device with L-CCO as the
tronic devices, the need for high-performance electrical energy storage positive electrode and active carbon (AC) as the negative electrode was
systems has become urgent. Supercapacitors have been paid much at­ fabricated. The device provides a substantial energy density of 58.7 Wh
tentions as one of the most important energy-storage systems because of kg 1 (64.5 μWh cm 2) at a power density of 800 W kg 1 (880 μW cm 2)
their high power density, excellent reversibility, fast charge–discharge and shows notable long-term stability (71.2% capacitance retention
rates and excellent cycling stability [1–5]. The electrochemical proper­ after 10,000 cycles within a broad operating potential window of 1.6 V).
ties of supercapacitors mainly depend on their electrode materials.
Various ternary transition metal oxides, including NiCo2O4 [6–8], 2. Experimental
CuCo2O4 [9–11], NiMn2O4 [12] and CoMoO4 [13], have been widely
investigated as pseudocapacitive electrode materials, owing to the 2.1. Fabrication of L-CCO electrodes
abundant oxidation states available for reversible redox reactions and
synergistic effects of multiple metal species compared with All chemical reagents used in this work were analytical grade and not
single-component metal oxides. Among these electrode materials, subjected to further purification. Before a typical procedure, a piece of
CuCo2O4 has attracted wide attentions as a considerate storage electrode Ni foam with dimensions of 1 � 4 cm2 was carefully cleaned in HCl
due to its greater electrochemical activity than single-component copper aqueous solution by sonication for 20 min and then sequentially rinsed
oxides or cobalt oxides [9,10,14–20]. Particularly, Cu, as an eco-friendly with acetone, ethanol and deionised (DI) water. Meanwhile, the growth
and cost-effective element, partial replacement of Co demonstrates to be solution was prepared by dissolving 2 mmol Co(NO3)2⋅6H2O, 1 mmol Cu
a better way of reduce the toxic characteristic and high-cost of cobalt (NO3)2⋅3H2O, 6 mmol NH4F and 20 mmol urea in 70 mL of DI water and
oxides. Nonetheless, CuCo2O4 still suffers from inferior conductivity and stirring for about 30 min at room temperature. Then, the resultant
insufficient concentrations of active sites, which adversely affects its reddish-brown solution was transferred to a 100 mL Teflon-lined
energy density and impedes its large-scale application. stainless steel autoclave, and the as-prepared cleaned Ni foam was
To overcome these challenges, various strategies have been pro­ immersed in the solution against the autoclave wall. Afterward, the
posed: (i) controlling the synthesis of CuCo2O4 electrodes to achieve autoclave was tightly sealed and maintained at 120 � C for 8 h. After
particular nanostructures (onion structures, nanosheets, double-shelled naturally cooling, the sample was removed, washed with DI water and
hollow spheres, etc.) with enhanced surface area [9,10,14]; (ii) ethanol several times and then dried in a vacuum oven at 60 � C for 5 h.
enwrapping CuCo2O4 particles with highly conductive materials (carbon Finally, three 1 � 1 cm2 Ni foam pieces cut from the same precursor were
nanotubes, carbon quantum dots, reduced graphenes, etc.) to improve annealed at 300 � C for 2 h under different mixed atmospheres of N2 and
their conductivity and inhibit their volume expansion [15–17]; and (iii) O2, denoted as low-crystalline CuCo2O4 (L-CCO, VN2:VO2 ¼ 9.5:0.5),
creating heteronanostructures by forming composites with other metal general-crystalline CuCo2O4 (G-CCO, VN2:VO2 ¼ 8:2) and high-
electrode materials (MnCo2O4, CuCo2S4, MnO2, etc.) to enrich the crystalline CuCo2O4 (H–CCO, VN2:VO2 ¼ 0:10). Then, the fabricated
electrochemical performance through synergistic effects [18,20,21]. electrodes were pressed under a pressure of 10 MPa for 1 min (Fig. S1).
However, despite all these recent progresses, the complex fabrication The mass loading of L-CCO, G-CCO and H–CCO was approximately 1.19,
processes and poor conductivity of the aforementioned materials 1.27 and 1.48 mg cm 2, respectively.
severely restrict their practical application. Therefore, intrinsically
improving the conductivity of CuCo2O4 electrodes are vital for fully 2.2. Fabrication asymmetric supercapacitor of L-CCO//AC
utilising their pseudocapacitance.
Notably, introduction of oxygen vacancies has been demonstrated as The asymmetric supercapacitor was assembled using L-CCO and AC
a convenient strategy to increase the electrical conductivity of metal as the positive and negative electrodes, respectively. The AC electrode
oxides and enlarge electroactive surface area exposed to the electrolyte. was prepared by scraping the homogeneous slurry of active carbon,
This approach is expected to be a facile, effective and reliable strategy to acetylene black and polyvinylidene fluoride (PVDF) in a weight ratio of
improve the capacitive properties of electrode materials [22–25]. 8:1:1 into a Ni foam current collector. In the approximated solid-state
Inhibiting crystallisation or using a reductant has recently been reported supercapacitor system, PVA/KOH gel was used as the electrolyte. In a
to be an effective strategy to induce oxygen vacancies. For instance, Cao typical gel electrolyte preparation, 6 g of PVA and 3.6 g of KOH were
and co-workers showed that introducing Pd2þ during a hydrothermal dissolved in 30 mL of DI water with stirring at 90 � C for 2 h until the gel
process can lead to poorly crystalline and oxygen-vacancy-rich Co3O4, became transparent. The as-obtained L-CCO and AC electrodes were
resulting in enhanced conductivity and redox reaction [26]. Lu et al. immersed in the PVA/KOH gel for approximately 3 min and then
developed an oxygen-vacancy strategy involving a simple phosphating assembled with a piece of cellulose paper between them in a sandwich
process, which substantially improved the electrochemical performance configuration; the resultant approximated solid-state asymmetric
of NiCo2O4 [25]. Despite substantial achievements having been made, to supercapacitor was used in subsequent testing.
our knowledge, there is no report on the effects of oxygen vacancies on
the electrochemical performance of CuCo2O4. Furthermore, the devel­ 2.3. Materials characterisation
opment of a low-cost and effective strategy to generate oxygen vacancies
remains a challenge. The morphologies and microstructures of samples were investigated
Herein, we report the fabrication of a mesoporous, low-crystalline using a field-emission scanning electron microscopy (FE-SEM, JEOL,
CuCo2O4 (L-CCO) nanoflower electrode with enriched oxygen va­ JSM-6701F) and transmission electron microscopy (TEM, JEOL, JEM-
cancies, which was synthesised by annealing precursor under a 2100) equipped with an energy dispersive X-ray spectrometer (EDS).
controllable N2:O2 volume ratio. The L-CCO, which has a chaotic lattice The crystallinity and phase purity of the obtained materials were
orientation and low crystallinity, functions as a shallow donor to examined by X-ray diffraction (XRD, M18XHF-SRA, Mac Science,
enhance the electrical conductivity as well as ensuring that sufficient Yokohama, Japan). The oxidation states and surface chemical compo­
electroactive sites are exposed to the electrolyte that can greatly sition of the products were characterised by X-ray photoelectron spec­
improve the kinetics of the redox reaction, resulting in excellent pseu­ troscopy (XPS, Thermo Electron, MultiLab2000). The specific surface
docapacitive performance [23,26–28]. Benefiting from the aforemen­ areas and pore structures of the electrode materials were calculated via
tioned advantages, the as-prepared L-CCO electrode exhibits a large the Brunauer–Emmett–Teller (BET) method using measurement data
specific capacitance of 1006 F g 1 at 1 A g 1 (1.2 F cm 2 at 1.2 mA from an ASAP 2460 (Micromeritics Instrument Corp.). The wetting

2
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

performances of the electrode materials were characterised using con­ electrochemical workstation (CHI 600E, Shanghai Chenhua) using a
tact angle measurements (Dataphysics OCT-20). three-electrode mode. Cyclic voltammetry (CV) and galvanostatic
charge–discharge (GCD) measurements as well as electrochemical
impedance spectroscopy (EIS) were carried out at ambient temperature
2.4. Electrochemical characterisation using a classical three-electrode configuration with 3 M KOH as the
electrolyte. The as-prepared electrodes were directly used as the
Electrochemical measurements were performed via an

Fig. 1. (a) Schematic illustration of the preparation of the L-CCO, G-CCO and H–CCO on Ni foam substrate. (b) XRD pattern of L-CCO, G-CCO and H–CCO. (c–e) SEM
images with different magnifications of the prepared L-CCO.

3
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

working electrode, a platinum plate acted as counter electrode and a shown in Fig. 1(c), L-CCO exhibits a flower-like structure uniformly
Hg/HgO electrode served as reference electrode in the three-electrode covering the Ni foam. High magnification SEM images (Fig. 1(d–e))
system. EIS measurements were performed in a frequency range of indicate that the sample consist of numerous 2D nanowires, and each
0.01 Hz–100 kHz at the open-circuit potential with 5 mV amplitude. The nanowire with a length of approximately 3 μm and an average diameter
formulas used to calculate the specific capacitance, energy density and of 25 nm. Furthermore, SEM images of as-prepared G-CCO and H–CCO
power density are given in the supporting information. electrodes are displayed in Fig. S2, revealing no obvious morphological
changes for the samples during different atmosphere processing. The
3. Results and discussion structures and morphologies of L-CCO, G-CCO and H–CCO were further
investigated by TEM. Fig. 2(a) presents TEM images of L-CCO at low
A schematic of the synthesis process of L-CCO, G-CCO and H–CCO on magnification, showing slender bent nanowires gathered together into
Ni foam is shown in Fig. 1(a). First, a thin layer of the precursors is nanoflowers, which is also consistent with the SEM results. Fig. 2(b–d)
formed on the Ni foam surface via a hydrothermal method. Then, the as- show the corresponding TEM images of L-CCO, G-CCO and H–CCO,
synthesised precursors are subsequent annealed in different calcination respectively. It is found that the nanowires of L-CCO G-CCO and H–CCO
conditions to form L-CCO, G-CCO or H–CCO, respectively (details in consist of a plentiful number of ultrasmall nanoparticles. Moreover,
Experimental Section). To elucidate the composition and crystalline there are numerous mesoporous structures in nanowires, which can be
structure of the products, XRD patterns were collected; the results are ascribed to the gaseous species releases from the CuCo-precursors during
illustrated in Fig. 1(b). The patterns of all samples show diffraction the thermal treatment process. Such the specific structure is favourable
peaks at 2θ angles of 18.9� , 31.0� , 36.7� , 38.4� , 44.7� , 55.3� , 59.2� and for electrolyte penetration when the materials are used as super­
65.1� , which can be perfectly indexed to the (111), (220), (311), (222), capacitor electrodes. High-resolution TEM (HRTEM) images of the three
(400), (422), (511) and (440) facets of spinel-structured CuCo2O4 samples (Fig. 2(e–g)) reveal that the H–CCO (Fig. 2(g)) is highly crys­
(JCPDS no. 76–1887). Notably, with increase in the volume ratio of N2 talline, with an interplanar spacing of 0.47 nm, which is consistent with
in the calcination condition, the intensity of the CuCo2O4 diffraction the CuCo2O4 (111) plane. Remarkably, the L-CCO sample with poor
peaks decreased, and simultaneously, the pattern showed wide diffrac­ polycrystallinity exhibits the smallest grain size, and no well-defined
tion peaks, suggesting L-CCO with weak crystallinity. Fig. 1(c–e) pre­ lattice fringes are observed after the hypoxic atmosphere treatment
sents different magnification SEM images of the as-prepared L-CCO. As (Fig. 2(e)). The small grain sizes and abundant amorphous regions

Fig. 2. Low TEM images of (a, b) L-CCO, (c) G-CCO and (d) H–CCO.High-resolution TEM image of (e) L-CCO, (f) G-CCO and (g) H–CCO. (h) EDS images of L-CCO. (i)
Nitrogen adsorption-desorption isotherms and (j) pore-size distributions of L-CCO, G-CCO and H–CCO.

4
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

enhance the immersion of the electrolyte ions and provide more sites for associated with lattice oxygen species (M O), oxygen defect sites and
the redox reaction, which is also consistent with the XRD analysis re­ hydroxyl groups (O–H), respectively [24–26]. The density of vacancy
sults. EDS was further used to verify the element composition, as shown defect sites in L-CCO (Fig. 3(d)) is clearly higher than those in G-CCO
in Fig. 2(h), demonstrating that elements Cu, Co and O are present. In and H–CCO [26,31]. The richness-enabled oxygen defects in the L-CCO
addition, Ni is also observed, originating from the Ni foam substrate. The electrode may result in enhanced energy-storage performance.
specific surface area (Fig. 2(i)) and pore size distribution (Fig. 2(j)) of To investigate the influence of oxygen vacancies on the electronic
the samples were investigated via the BET and Barrett–Joyner–Halenda structure of CuCo2O4, we conducted first-principles simulations to study
(BJH) methods, respectively. The corresponding surface area of L-CCO the electronic structures of perfect CuCo2O4 and CuCo2O4 with O-atom
(62.9 m2 g 1) is distinctly larger than those of G-CCO (52.8 m2 g 1) and vacancies (V–CCO); the results are displayed in Fig. 4. Fig. 4(a and b)
H–CCO (51.4 m2 g 1). Meanwhile, the average pore diameter distribu­ shows the atomic structure models of pristine CuCo2O4 and V–CCO. The
tion for L-CCO (7.9 nm) is smaller than those for G-CCO (9.3 nm) and band gaps of CuCo2O4 and V–CCO are shown in Fig. 4(c and d). Two
H–CCO (10.7 nm), which is consistent with the TEM results. The changes impurity bands appear at the Fermi energy of V–CCO; these bands can
in the surface areas and pore diameters can be attributed to the decrease possibly enhance the electron conduction from the valence band to the
in crystallinity, which is unfavourable for grain growth, resulting in a conduction band, confirming that the introduction of oxygen vacancies
smaller grain size and larger specific surface area. This observation is effectively enhances the electrical conductivity of CuCo2O4. Fig. 4(e and
also consistent with those in previous works for low crystallinity mate­ f) shows the calculated density of states of CuCo2O4 and V–CCO. Obvi­
rials. Obviously, the introduction of oxygen vacancies positively affects ously, the Fermi level shifts towards the conduction band after the
the surface structure of the samples, providing many electrochemically introduction of oxygen vacancies, further confirming that our strategy
active sites for redox reaction and facilitating electrolyte-ion penetration can effectively enhance the conductivity of the V–CCO.
into the surface of the active materials. The electrochemical behavior of the three samples were charac­
XPS analysis was carried out to further investigate the chemical terised using a three-electrode electrochemical system with aqueous
composition and oxidation state of the as-prepared L-CCO, G-CCO and KOH solution (3 M) as the electrolyte. Fig. 5(a) shows a comparison of
H–CCO. All Co 2p emission spectra (Fig. 3(a)) show two spin–orbit the CV curves of L-CCO, G-CCO and H–CCO at a scan rate of 10 mV s 1 in
doublets located at approximately 796 and 780 eV. With decreasing the potential range from 0 to 0.6 V (vs. Hg/HgO). Notably, the enclosed
crystallinity, the intensities of the peaks at 780.8 and 796.3 eV, assigned area in the cyclic voltammogram of L-CCO is larger than those in the
to Co2þ, increase substantially, indicating that more Co3þ ions were voltammograms of G-CCO and H–CCO, indicating that the L-CCO ex­
transformed into Co2þ in the sample of L-CCO and that oxygen vacancies hibits the highest specific capacitance among the three samples. In
were introduced [25,26,29]. Similar changes are observed in the Cu 2p addition, with the conversion between different cobalt and copper
XPS spectra (Fig. 3(b)). Two shake-up satellite peaks are easily observed oxidation states, the CV curves of these three electrodes exhibit a pair of
at 941.6 and 961.8 eV, confirming the presence of Cu2þ. In the spectra of well-defined redox peaks, signifying typical battery-type electro­
samples annealed under a hypoxic atmosphere (L-CCO), the peaks of Cu chemical behavior and mainly based on the following redox reactions
2p3/2 and Cu 2p1/2 shift towards lower binding energies, as expected, [9,16]:
indicating that more Cuþ is formed in the L-CCO sample to compensate
CuCo2 O4 þ H2 O þ e ↔ 2CoOOH þ CuOH (1)
for the oxygen vacancies [16,30]. Fig. 3(c) shows the O 1s spectra of
L-CCO, G-CCO and H–CCO. The spectra of all samples show three
CoOOH þ OH ↔ CoO2 þ H2 O þ e (2)
characteristic peaks centred at 529.4, 531.2 and 532.5 eV, which are

Fig. 3. High-resolution XPS spectra of (a) Co 2p, (b) Cu 2p, (c) O 1s for L-CCO. (d) The concentration of Co2þ and oxygen defect calculated from XPS data.

5
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

Fig. 4. The atomic structure of (a) CuCo2O4 and (b) V–CCO. Density functional theory (DFT) calculation total band structures of (c) CuCo2O4 and (d) V–CCO, the
density of states (DOS) of (e) CuCo2O4 and (f) V–CCO.

CuOH þ OH ↔ CuðOHÞ2 þ e (3) exhibits two distinct parts including a small semicircle in the high fre­
quency region and a straight line in the low frequency region, corre­
GCD tests of the three samples were further conducted at a current sponding to the charge-transfer reaction and ions diffusion, respectively
density of 1 A g 1, as shown in Fig. 5(b). As expected, the charge and [32–34]. The Z-view software has been used to clarify the impedance
discharge times of the L-CCO electrode were relatively longer than those according to an equivalent circuit (see the inset in Fig. 5(d)). The EIS
of the G-CCO and H–CCO electrodes, which is consistent with the CV data can be fitted by a pseudocapacitance (CPE2) from redox operation
results. Collectively, the results indicate that the L-CCO electrode ex­ of composites, bulk solution resistance (Rs), a charge transfer resistance
hibits superior electrochemical performance compared with the G-CCO (Rct), and a constant phase element (CPE1) [35–37]. After fitting the EIS
and H–CCO electrodes, implying that the L-CCO electrode yields the spectra, both Rs (1.57 Ω) and Rct (0.40 Ω) of the L-CCO system are the
highest specific capacitance. This superior performance is attributed to smallest values as compared to those of the G-CCO (Rs: 1.72 Ω; Rct: 0.75
the presence of richness-enabled oxygen defects on the surface, which Ω) and H–CCO (Rs: 1.76 Ω; Rct: 0.80 Ω) systems. This result reveals that
can provide more active sites and lead to fast transmission of electrons the introduction of oxygen vacancies can indeed effectively improve the
and ions. The poor crystallinity and oxygen vacancies are also beneficial conductivity of the electrodes and boost their redox reaction kinetics.
to the wetting ability, increasing the active substance contact with the The capacitive performances of the L-CCO electrodes were compre­
electrolyte. The static contact angle between the three samples and 3 M hensively measured and evaluated. The CV curves of the L-CCO elec­
KOH electrolyte was recorded; the results (Fig. S3) show that the static trode (Fig. 5(e)) exhibit unchanging shapes as the scan rate increases
contact angle increases from 48.6� for L-CCO to 55.2� for H–CCO. from 5 to 100 mV s 1, demonstrating its excellent pseudocapacitive
To further evaluate the influence of the oxygen defects on the elec­ behavior with fast and reversible charge storage capability in the po­
trochemical properties of the CuCo2O4 electrode, the capacitances of L- tential window from 0 to 0.6 V. Furthermore, the GCD measurements at
CCO, G-CCO and H–CCO electrodes as a function of current density various current densities are displayed in Fig. 5(f). All the curves at
(ranging from 1 to 20 A g 1 or 1.2–24 mA cm 2) are compared in Fig. 5 various current densities are nonlinear and almost symmetric, indicating
(c) and Fig. S4. The L-CCO electrodes exhibit the highest specific ca­ an enhanced pseudocapacitive properties and a high coulombic effi­
pacity of 1006 F g 1 (or areal capacitance of 1.2 F cm 2 and volumetric ciency. The CV and GCD curves of G-CCO and H–CCO are also shown in
capacitance of 141 F cm 3) than the G-CCO and H–CCO electrodes, Fig. S5.
revealing that the oxygen-vacancy functionalization profoundly affects For supercapacitor practical application, its cycling performance is a
the charge–discharge properties of the L–CCO–based electrodes. More­ significant factor. Therefore, the cyclic stability of the as-prepared
over, the L-CCO electrode shows good reversible capability, with 69.4% electrodes was investigated using CV measurements at 50 mV s 1 as
retention of its initial specific capacitance when the current density is shown in Fig. 5(g). After 10,000 cycles, the specific capacitance of the
increased 20-fold. electrodes maintains 85.5% of initial capacitance. Moreover, the CV and
In order to investigate the influence of introduced oxygen vacancies EIS curves show unobvious variations even after prolonged 10,000 scans
on the electrode kinetics, EIS measurements were carried out at room (Fig. 5(h) and Fig. S6), suggesting that the fabricated L-CCO electrodes
temperature; the results are shown in Fig. 5(d). Each Nyquist plot with great redox reversibility can undergo long-term cycling tests. In

6
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

Fig. 5. Electrochemical performance of the H–CCO, G-CCO and L-CCO electrode: (a) CV curves at a scan rate of 10 mV s 1, (b) GCD tests at a current densities of 1 A
g 1, (c) specific capacitance at different current densities, (d) EIS spectra. Electrochemical performance of L-CCO electrode: (e) CV curves at different scan rates, (f)
GCD tests at various current densities, (g) cycling stability at a current densities of 50 mV s 1, and (h) the first and last 10,000 cycle plot. (i) Typical SEM image of L-
CCO after 10,000 cycles. (j) Schematic illustration displaying the merits of the L-CCO electrode for energy storage.

addition, the SEM morphology of L-CCO electrode cycled after 10,000 the grains in the nanowires of L-CCO facilitate electrolyte penetration
times is also shown in Fig. 5 (i). Clearly, after the long-term cycling, the into the inner part of the nanowires, making full use of the active elec­
overall L-CCO architecture is evidently preserved with little structural trode. Third, the direct growth of L-CCO on a highly conductive Ni foam,
deformation, suggesting its robust mechanical stability. Compared with without polymer binders and conductive agents, reduces the contact
other reported CuCo2O4, copper or cobalt oxide/hydroxide based com­ resistance and shortens the charge transmission route. Finally, the
posites and other bimetallic oxide, the as-synthesised L-CCO electrode unique structure with countless nanowires is firmly anchored onto the
shows superior capacitance and outstanding cycling performance central nanospheres, making them completely accessible to the elec­
(Table S1). trolyte and providing a fast ion-transport channel. Additionally, the
The superior electrochemical performance of L-CCO is attributable to open space of nanowires and central nanospheres can accommodate a
the introduction of oxygen vacancies and its unique nanostructure, as volume change, thereby suppressing expansion or contraction during
shown in Fig. 5(j). Most importantly, oxygen vacancies substantially repeated charging and discharging processes.
boost the reactive active sites towards the electrolyte and improve To evaluate the practical applications of the L-CCO electrode, addi­
electrical conductivity, resulting in intensive reaction kinetics and tional electrochemical tests were conducted. An approximated solid-
reversible redox reactions. Second, the abundant mesopores between state asymmetric supercapacitor device with a sandwich-like structure

7
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

was fabricated, where AC was used as the anode (potential windows Notably, the total capacitance displayed good cycling stability, with a
from 1 to 0 V, Fig. 6(a) and Fig. S7). To achieve high charge storage capacitance retention of 71.2%, indicating excellent electrochemical
performance, the charge between the positive and negative electrodes stability of the asymmetric supercapacitor device. Gravimetric power
should be balanced; the calculated optimal mass ratio between the L- density and energy density are two key parameters that should be
CCO cathode and AC anode was 2.8 (details in Supporting Information). considered when assessing the practical applicability of electrochemical
After assembling the device, we evaluated the stable electrochemical supercapacitors (Fig. 6(h)). Remarkably, the device synthesised in our
window of the L-CCO electrode to avoid damaging the device. According work reaches a maximum energy density of 58.7 Wh kg 1 at a power
to the CV tests, the operating cell voltage of the as-fabricated L-CCO density of 800 W kg 1 (or 64.5 μWh cm 2 at 880 μW cm 2, Fig. S8(b)),
device can be extended to as high as 1.6 V without obvious polarisation, which is much higher than the energy densities of recently reported
as shown in Fig. 6(b). The CV curves of the as-fabricated L-CCO device supercapacitor devices such as NiCo2O4//AC (23.6 Wh kg 1 at 585.5 W
were recorded at different scan rates from 5 to 200 mV s 1, as illustrated kg 1) [38], ZnCo2O4//AC (39.4 Wh kg 1 at 650 W kg 1) [39], NiC­
in Fig. 6(c). The shape of the CV curves of the device was well main­ oMn–OH//RGO (42.8 Wh kg 1 at 749 W kg 1) [40],
tained even at higher scan rates, demonstrating good reversibility for Ni0⋅67Co0⋅33MoO4//RGO (25.7 Wh kg 1 at 775 W kg 1) [41],
power devices. To further evaluate the electrochemical performance of FeCo2O4//FeCo2O4 (30.9 Wh kg 1 at 1551 W kg 1) [42],
the device, GCD measurements were conducted at different current CuCo2O4//rGO (45.2 Wh kg 1 at 1500 W kg 1) [43] and NiCo2S4@­
densities (Fig. 6(d)). As shown in Fig. 6(e) and Fig. S8(a), the as- graphene//porous carbon (43.4 Wh kg 1 at 254.3 W kg 1) [44].
fabricated L-CCO device delivers a specific capacitance ranging from Notably, the energy density remained 25.3 Wh kg 1 as the power den­
165 to 71.3 F g 1 (181.5–78.4 mF cm 2) with an increase in current sity increased to 16,000 W kg 1. To demonstrate the practical applica­
density from 1 to 20 A g 1 (1.5–31 mA cm 2). As revealed in Fig. 6(f) tion of the fabricated electrode in energy storage, a commercial
and (g), the cycle performance of the device was further measured by light-emitting diode is lighted by two L-CCO devices connected in se­
GCD measurements at a current density of 10 A g 1 for 10,000 cycles. ries (Fig. 6(i)), demonstrating its potential suitability for energy-storage

Fig. 6. (a) CV curves of AC and L-CCO from 1.0 to 0 V and 0–0.6 V, respectively, in a three-electrode system. (b) CV curves of as-assembled L-CCO//AC device in
different operation voltages at a scan rate of 20 mV s 1. (c) CV curves, (d) GCD curves, (e) capacitance behaviors versus current density of L-CCO//AC device. (f)
Cycling stability of the L-CCO//AC device at a current density of 10 A g 1 during 10,000 cycles. (g) GCD curves of the 1st, 10000th cycle for the L-CCO//AC device.
(h) Ragone plot of the ASC device. (i) Image of the red LED powered by L-CCO//AC device. (For interpretation of the references to colour in this figure legend, the
reader is referred to the Web version of this article).

8
Y. Feng et al. Journal of Power Sources 458 (2020) 228005

applications. [6] C. Guan, X. Liu, W. Ren, X. Li, C. Cheng, J. Wang, Adv. Energy Mater. 7 (2017),
1602391.
[7] B. Liu, J. Hou, T. Zhang, C. Xu, H. Liu, J. Mater. Chem. 7 (2019) 16222–16230.
4. Conclusions [8] Z. Qu, M. Shi, H. Wu, Y. Liu, J. Jiang, C. Yan, J. Power Sources 410–411 (2019)
179–187.
[9] A.A. Ensafi, S.E. Moosavifard, B. Rezaei, S.K. Kaverlavani, J. Mater. Chem. 6 (2018)
In summary, a mesoporous low-crystalline CuCo2O4 with enriched 10497–10506.
oxygen vacancies was successfully synthesised via a facile in situ hy­ [10] A.T. Aqueel Ahmed, B. Hou, H.S. Chavan, Y. Jo, S. Cho, J. Kim, S.M. Pawar, S. Cha,
drothermal method followed by thermal treatment in a hypoxic atmo­ A.I. Inamdar, H. Kim, H. Im, Small 14 (2018) 1800742.
[11] M.M. Vadiyar, X. Liu, Z. Ye, J. Power Sources 415 (2019) 154–164.
sphere. L-CCO shows a specific capacitance of 1006 F g 1 at 1 A g 1,
[12] W. Zuo, C. Xie, P. Xu, Y. Li, J. Liu, Adv. Mater. 29 (2017), 1703463.
which is much higher than those of G-CCO and H–CCO. This superior [13] Q. Li, Y. Li, J. Zhao, S. Zhao, J. Zhou, C. Chen, K. Tao, R. Liu, L. Han, J. Power
pseudocapacitive performance is attributed to the low crystallinity of Sources 430 (2019) 51–59.
the L-CCO electrode with a mostly amorphous nanostructure and [14] S.K. Kaverlavani, S.E. Moosavifard, A. Bakouei, Chem. Commun. 53 (2017)
1052–1055.
abundant oxygen vacancies, which can provide more reaction sites and [15] X. Wang, Y. Li, T. Jin, J. Meng, L. Jiao, M. Zhu, J. Chen, Nano Lett. 17 (2017)
fast ion intercalation, resulting in fast faradaic redox reactions and 7989–7994.
outstanding rate capability. More importantly, an asymmetric super­ [16] G. Wei, X. Zhao, K. Du, Y. Huang, C. An, S. Qiu, M. Liu, S. Yao, Y. Wu, Electrochim.
Acta 283 (2018) 248–259.
capacitor device assembled with L-CCO and AC as the positive and [17] P. Liang, F. Wang, Z.-A. Hu, Chem. Eng. J. 350 (2018) 627–636.
negative electrodes delivered a maximum energy density of 58.7 Wh [18] S. Liu, K.S. Hui, K.N. Hui, J.M. Yun, K.H. Kim, J. Mater. Chem. 4 (2016)
kg 1 at a power density of 800 W kg 1. The device demonstrates 8061–8071.
[19] J. Lin, H. Jia, H. Liang, S. Chen, Y. Cai, J. Qi, C. Qu, J. Cao, W. Fei, J. Feng, Chem.
outstanding cycling stability in a wide potential window of 1.6 V with Eng. J. 336 (2018) 562–569.
71.2% capacitance retention even after 10,000 cycles. Our work reveals [20] H. Ge, C. Wang, L. Yin, J. Mater. Chem. 3 (2015) 17359–17368.
that the oxygen vacancies can play an important role in enhancing [21] X. Xu, Y. Liu, P. Dong, P.M. Ajayan, J. Shen, M. Ye, J. Power Sources 400 (2018)
96–103.
electrochemical performance, which is also expected to be effective [22] H.-S. Kim, J.B. Cook, H. Lin, J.S. Ko, S.H. Tolbert, V. Ozolins, B. Dunn, Nat. Mater.
when applied to other transition metal oxides. Additionally, our prep­ 16 (2017) 454–460.
aration of asymmetric supercapacitors with high power density and [23] G. Zhang, T. Xiong, M. Yan, L. He, X. Liao, C. He, C. Yin, H. Zhang, L. Mai,
Nanomater. Energy 49 (2018) 555–563.
energy density will enrich the next generation of energy-storage devices.
[24] T. Zhai, L. Wan, S. Sun, Q. Chen, J. Sun, Q. Xia, H. Xia, Adv. Mater. 29 (2017),
1604167.
[25] Y. Zeng, Z. Lai, Y. Han, H. Zhang, S. Xie, X. Lu, Adv. Mater. 30 (2018), 1802396.
Declaration of competing interest [26] J. Hao, S. Peng, H. Li, S. Dang, T. Qin, Y. Wen, J. Huang, F. Ma, D. Gao, F. Li,
G. Cao, J. Mater. Chem. 6 (2018) 16094–16100.
The authors declare that they have no known competing financial [27] D. Yan, W. Wang, X. Luo, C. Chen, Y. Zeng, Z. Zhu, Chem. Eng. J. 334 (2018)
864–872.
interests or personal relationships that could have appeared to influence [28] G. Zhu, C. Xi, M. Shen, C. Bao, J. Zhu, ACS Appl. Mater. Interfaces 6 (2014)
the work reported in this paper. 17208–17214.
[29] S. Yang, Y. Liu, Y. Hao, X. Yang, W.A. Goddard III, X.L. Zhang, B. Cao, Adv. Sci. 5
(2018), 1700659.
Acknowledgements [30] C. Jin, Y. Cui, G. Zhang, W. Luo, Y. Liu, Y. Sun, Z. Tian, W. Zheng, Chem. Eng. J.
343 (2018) 331–339.
This work was financially supported by the National Natural Science [31] S. Hong, R.M. Doughty, F.E. Osterloh, J.V. Zaikina, J. Mater. Chem. 7 (2019)
12303–12316.
Foundation of China (11604395), the Key Technologies R&D Program of [32] J.A. Syed, J. Ma, B. Zhu, S. Tang, X. Meng, Adv. Energy Mater. 7 (2017), 1701228.
Henan Province (192102210195), the Program for Science and Tech­ [33] F.B. Ajdari, E. Kowsari, A. Ehsani, J. Solid State Chem. 265 (2018) 155–166.
nology Innovation Talents in Universities of Henan Province (18HAS­ [34] F.B. Ajdari, E. Kowsari, A. Ehsani, M. Schorowski, T. Ameri, Electrochim. Acta 292
(2018) 789–804.
TIT032) and Chongqing Key Laboratory for Advanced Materials &
[35] A. Ehsani, M. Bigdeloo, M.Y. Ansari, B. Mirtamizdoust, A.A. Heidari, M. Hadi, H.
Technologies of Clean Energies (Grant No. JJNY201901). M. Shiri, Bull. Chem. Soc. Jpn. 91 (2018) 617–622.
[36] H.M. Shiri, A. Ehsani, M.J. Khales, J. Colloid Interface Sci. 505 (2017) 940–946.
[37] M. Naseri, L. Fotouhi, A. Ehsani, H.M. Shiri, J. Colloid Interface Sci. 484 (2016)
Appendix A. Supplementary data 308–313.
[38] D. Sun, Y. Li, X. Cheng, H. Shi, S. Jaffer, K. Wang, X. Liu, J. Lu, Y. Zhang,
Supplementary data to this article can be found online at https://doi. Electrochim. Acta 279 (2018) 269–278.
[39] Q. Wang, Y. Zhu, J. Xue, X. Zhao, Z. Guo, C. Wang, ACS Appl. Mater. Interfaces 8
org/10.1016/j.jpowsour.2020.228005.
(2016) 17226–17232.
[40] F. Lai, Y.-E. Miao, L. Zuo, H. Lu, Y. Huang, T. Liu, Small 12 (2016) 3235–3244.
References [41] H.C. Chen, Y. Qin, H. Cao, X. Song, C. Huang, H. Feng, X.S. Zhao, Energy Storage
Mater 17 (2019) 194–203.
[42] B. Zhu, S. Tang, S. Vongehr, H. Xie, J. Zhu, X. Meng, Chem. Commun. 52 (2016)
[1] Z. Yu, L. Tetard, L. Zhai, J. Thomas, Energy Environ. Sci. 8 (2015) 702–730.
2624–2627.
[2] Y. Shao, M.F. El-Kady, J. Sun, Y. Li, Q. Zhang, M. Zhu, H. Wang, B. Dunn, R.
[43] S.K. Kaverlavani, S.E. Moosavifard, A. Bakouei, J. Mater. Chem. 5 (2017)
B. Kaner, Chem. Rev. 118 (2018) 9233–9280.
14301–14309.
[3] P. Simon, Y. Gogotsi, B. Dunn, Science 343 (2014) 1210–1211.
[44] F. Yu, Z. Chang, X. Yuan, F. Wang, Y. Zhu, L. Fu, Y. Chen, H. Wang, Y. Wu, W. Li,
[4] Y. Jiang, J. Liu, Energy Environ. Mater. 2 (2019) 30–37.
J. Mater. Chem. 6 (2018) 5856–5861.
[5] X. Chen, R. Paul, L. Dai, Natl. Sci. Rev. 4 (2017) 453–489.

You might also like