Moradi 2023

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Energy Storage 71 (2023) 108177

Contents lists available at ScienceDirect

Journal of Energy Storage


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

Research papers

Electrosynthesis of Co–Mn layered-double-hydroxide as a precursor for


Co-Mn-MOFs and subsequent electrochemical sulfurization for
supercapacitor application
Mahdi Moradi a, Abbas Afkhami a, b, *, Tayyebeh Madrakian a, **, Hamid Reza Moazami c
a
Faculty of Chemistry, Bu Ali Sina University, Hamedan, Iran
b
D-8 International University, Hamedan, Iran
c
Nuclear Fuel Cycle Reseach School, Nuclear Science and Technology Research Institute (NSTRI), Tehran, Iran

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

Keywords: Today, researchers are focusing on improving the electrochemical efficiency of supercapacitors by designing and
Electrosynthesis evolving synthesis routes for battery-type hierarchical materials. This study proposes a green method for creating
MOF derived high-efficiency positive electrode materials with a mesoporous heterostructure on Ni foam (NF) using a metal-
Electrochemical transformation
organic framework (MOF)-derived approach. A binder-free electrosynthesis technique was utilized to grow
Co-Mn− S
Battery-type electrode materials
CoMn-layered double hydroxide (CoMn-LDH) and exchange ions to obtain a porous metal precursor, namely
Asymmetric supercapacitors binary Co-Mn-MOF. The MOF arrays were then electrochemically converted to CoMn-sulfide (CMS), which ex­
hibits excellent conductivity and a mesoporous structure, enabling it to serve as a 3D continuous network for ion
and electron conduction in energy storage. The synthesized sample was characterized using various techniques,
including field-emission scanning electron microscopy (FESEM), elemental mapping, high-resolution trans­
mission electron microscopy (HR-TEM), energy-dispersive X-ray analysis (EDX), Brunauer-Emmett-Teller (BET)
surface area analysis, X-ray photoelectron spectroscopy (XPS), and X-ray diffraction (XRD). Among the prepared
electrode samples, CoMn− S with an Mn/Co feeding ratio of 1:2 demonstrated outstanding electrochemical
properties. Based on this platform, CMS was applied and validated as a positive electrode in supercapacitors,
exhibiting a high specific capacity of 1091C g− 1 at 1 A g− 1 in a three-electrode configuration and remarkable
cycling life (85 % capacitance retention over 7000 cycles at 25 A g− 1). Furthermore, an assembled asymmetric
supercapacitor (ASC) device using CMS as the cathode and AC as the anode demonstrated satisfactory electro­
chemical performance. The CMS//AC device delivers an energy density of 84 Wh kg− 1 at a high power density of
1191.4 W kg− 1 and maintains stable electrochemical stabilities (92 % capacitance retention even after 7000
cycles). This synthesis approach opens up a new avenue for developing binder-free electrodes with hierarchical
structures using metal sulfides, thereby enhancing the electrochemical performance of hybrid supercapacitors.

1. Introduction asymmetric or hybrid SCs using adaptable materials. Hybrid SCs


combine carbon materials with non-faradaic electric double-layer
Electrical energy storage systems (EESs) are a promising solution to capacitor (EDLC) charge storage mechanisms and pseudocapacitive/
address the challenges of energy demand, greenhouse emissions, and battery-type materials with a Faradaic mechanism in a single-cell
global warming [1]. Supercapacitors (SCs) have gained attention in configuration [8,9]. Battery-type materials, such as transition metal
R&D due to their high-power density, long cycle life, and excellent oxides, sulfides, and selenides, store charge through faradic reactions,
discharge time [2–4]. However, their lower energy density compared to with specific capacitance values varying with the voltage window and,
rechargeable batteries hinders their practical application [5–7]. One C/g or mAh/g metric [10–12].
effective approach to achieving high energy density while maintaining Among battery-type materials, transition metal sulfides (TMSs) like
power density is improving the potential window by assembling CoS2, NiSx, MnS2, VS2, and NiCo2S4 have garnered significant

* Correspondence to: A. Afkhami, Faculty of Chemistry, Bu Ali Sina University, Hamedan, Iran.
** Corresponding author.
E-mail addresses: afkhami@basu.ac.ir (A. Afkhami), madrakian@basu.ac.ir (T. Madrakian).

https://doi.org/10.1016/j.est.2023.108177
Received 15 March 2023; Received in revised form 8 June 2023; Accepted 26 June 2023
Available online 10 July 2023
2352-152X/© 2023 Elsevier Ltd. All rights reserved.
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 1. Schematic illustration of the fabrication process of the nanostructured CoMn-S grown uniformly on Ni foam.

Fig. 2. Current-time (Chronoamperometry) curve of the Electrodeposition of Co-Mn-LDH at − 1.0 V (vs. Ag/AgCl), (a) Typical cyclic voltammograms (CV) curves
during the EC-MOF process (initial 250 CV cycles, scan rate: 0.2 V s− 1) (b).

attention due to their properties. Sulfides have a short band gap that and manganese, enhance conductivity and capacity. Cobalt and man­
increases electrical conductivity, facilitates electron transfer, and en­ ganese sulfide compounds are cost-effective, abundant in nature, envi­
hances electrochemical efficiency. The low electronegativity of sulfur ronmentally friendly, and offer high theoretical capacity [27].
allows for structural expansion, improving mechanical flexibility and Metal-organic frameworks (MOFs) with flexible structures composed
cycle life [13,14]. Mixed metal sulfides exhibit higher capacity than of metal ions and organic linkers are suggested as templates and pre­
monometallic sulfides due to various oxidation states and synergistic cursors with high surface area for active materials with desirable ar­
effects, resulting in enhanced electrical conductivity and electro­ chitectures and adjustable compositions. Raw MOFs can be converted
chemical activity [15,16]. into carbon materials, metal oxides, or metal sulfides with desired
Various mixed transition metal sulfides, including CuCo2S4 [17,18], structures using simple, controllable, and effective methods. Conven­
ZnCo2S4 [19,20], FeCo2S4 [21,22], MnCo2S4 [23–25], and NiCo2S4 [26] tional synthesis methods for MOF-derived materials often rely on chal­
have been extensively explored as supercapacitor electrode materials. lenging experimental conditions like hydrothermal synthesis and
CoMn-S (CMS) is a promising material for supercapacitor electrodes. pyrolysis. However, these methods have drawbacks, including high
The high oxidation potential of cobalt, along with the increased elec­ energy and time requirements and rapid collapse of the primary MOF
trons supplied by manganese species and charge transfer between cobalt structure. Thus, developing a mild, green, and sustainable method to

2
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

preparation of metal‑sulfuric compounds in the development of high-


efficiency energy storage devices.

2. Experimental section

2.1. Synthesis

2.1.1. Chemicals &materials


Manganese(II) nitrate tetrahydrate, Mn(NO3)2.4H2O, cobalt nitrate
hexahydrate, Co(NO3)2.6H2O, benzene-1,3,5-tricarboxylic acid (BTC),
Thiourea, ethanol, hydrochloric acid, HCl (37 %), Potassium nitrate,
KNO3, Potassium chloride, KCl, and Potassium hydroxide, KOH, were
purchased from Merck Company, Ni foam (NF, 0.3 mm thickness),
polytetrafluoroethylene (PTFE, 60 wt% in water), activated carbon
(AC), and carbon black were obtained from Redoxkala, Tehran, Iran,
deionized water (DIW).

2.1.2. Electrosynthesis of MnCo- LDH


Before each synthesis, the NF substrate underwent a pretreatment
process involving sonication with 3 M HCl, deionized water (DIW), and
ethanol for 15 min in the given order. This process aimed to eliminate
organic impurities and nickel oxides formed on the surface. Hierarchical
MnCo-LDH nanosheets were synthesized on the NF substrate using an
electrosynthesis strategy. An electrochemical cell was constructed with
an aqueous solution containing Co(NO3)2⋅6H2O, Mn(NO3)2⋅4H2O, and
KNO3, and a three-electrode arrangement that consisted of the NF sub­
strate, platinum plate, and Ag/AgCl as the working, counter, and
reference electrodes, respectively. Electrodes with different Co:Mn ra­
tios (3:0, 1:2, 1:1, 1:2, and 0:3) were prepared, maintaining a total salt
concentration of 0.05 M and a KNO3 concentration of 0.3 M. CoMn-LDHs
were synthesized using Chronoamperometry at a constant potential of
− 1.0 V vs. Ag/AgCl for a duration of 400 s. The electrodes were washed
with DIW after deposition and dried at room temperature.

2.1.3. Synthesis of Mn–Co MOF


For the preparation of MnCo-MOF, the LDHs/NF samples were sub­
Fig. 3. XRD patterns of the (a) CoMn-LDH, (b) CoMn-MOF, (c) CoMn-S. merged in a 0.15 M BTC ethanol solution consisting of 0.3 g BTC, 2 mL
DIW, and 7 mL absolute ethanol. The immersion time was optimized to
convert MOFs into active nanomaterials with high surface area for en­ 20 min to ensure the complete conversion of MnCo-LDHs to MnCo-BTC
ergy storage applications with a controllable conversion rate and degree MOFs. Subsequently, the prepared electrode was rinsed with DIW to
is of great importance. Electrosynthesis of MOF layers on NF substrates eliminate any residual unreacted BTC from the NF surface.
as current collectors enhances the electrical conductivity of the prepared
electrode by eliminating the need for an organic binder. This improve­ 2.1.4. Electrosynthesis of manganese cobalt sulfides (MCS)
ment can lead to increased accuracy and repeatability of the prepared The electrochemical conversion of MOF films into hierarchical
electrode. MnCo-S was carried out using cyclic voltammetry in a typical three-
In this research, a binder-free electrosynthesis approach is utilized to electrode system. The MnCo-MOF on NF served as the working elec­
grow CoMn-layered double hydroxide (CoMn-LDH) as a porous metal trode, while an Ag/AgCl/saturated KCl electrode and a platinum plate
precursor for the preparation of binary Co-Mn-MOF through facile ion were used as the reference and counter electrodes, respectively. The
exchange. A controlled electrochemical method is then introduced to electrolyte for the sulfurization step consisted of 0.5 M Thiourea and 1 M
convert the MOF into 3D hierarchically porous amorphous CoMn-sulfide KCl. Prior to the experiment, the electrolyte solution was purged with N2
(CMS). Optimization of the Co/Mn ratios and reaction conditions en­ gas and stirred for at least 15 min to remove dissolved oxygen. Cyclic
sures the formation of the desired hierarchical architecture, maximizing voltammetry cycles were performed between − 1.5 and +0.2 V (vs. Ag/
the contact surface between the active material and electrolyte, and AgCl/saturated KCl) at a scan rate of 0.2 Vs− 1 for a total of 750 cycles to
achieving high electrical conductivity during synthesis. The synthesized complete the synthesis. Following the electrochemical conversion, the
CMS exhibits outstanding performance when used as an electrode in MnCo-S working electrode was rinsed with DIW and dried in a vacuum
supercapacitors, demonstrating a capacitance of 1091 Cg− 1 at 1 Ag− 1 oven at 60 ◦ C.
and capacity retention of 58.2 % from 1 to 20 Ag− 1. The loading mass of the active material on NF was calculated using
An Asymmetric Supercapacitor (ASC) device (with operating po­ the equation mn − mn-1, where n is the synthesis step. The mass loading
tential = 1.6 V) is assembled using CMS as the positive electrode, acti­ after coating with CoMn-LDH ranged from 2.1 to 2.3 mg cm− 2. After
vated carbon (AC) as the negative electrode, and filter paper as the MOF formation, the mass loading on the NF was increased to about 3 to
separator, with KOH (3 M) as the electrolyte. This device exhibits a high 3.5 mg cm− 2. However, during the EC-MOF process, there was a mass
energy density of 84 W h kg− 1 and an excellent power density of 1191.4 loss, resulting in a mass loading of 1.3–1.5 mg cm− 2. This mass loss can
W kg− 1 at 1 A g− 1, along with remarkable cycling stability (92 % be attributed to the larger molecular mass of released BTC (210.14 g
capacitance retention after 7000 cycles). The results of this research mol− 1) compared to that of added S2− ions (32 g mol− 1), and the
confirm the effectiveness of the designed route for the hierarchical shedding amount of Co and Mn from the structure during the electro­
chemical conversion.

3
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 4. FE-SEM images of the (a) CoMn-LDH, (b) CoMn-MOF, (c) CoMn-S.

2.1.5. Preparation of activated carbon (AC) electrode material, whereas the counter and reference electrodes were platinum
The negative electrode was prepared by dispersing an AC, carbon plate (1 × 2 cm) and Ag/AgCl, respectively. Cyclic voltammetry (CV)
black, and PTFE, (8:1:1 mass ratio) in deionized water, then coating NF data were collected with a potential window of − 0.1 to 0.55 V and a scan
(1 × 1 × 0.03 cm) with the prepared mixture and drying the prepared rate of 1 to 75 mV s− 1. The galvanostatic charge-discharge (GCD) curves
electrode at 80 ◦ C for 12 h. were measured at current densities from 1 to 25 A g− 1. Based on the
charge-discharge curves, the specific capacity, energy density, and
2.2. Characterization power density is calculated by Eqs. (1)–(3).
∫vb
2.2.1. Materials characterization Cs =
1
I(V)dV (1)
The crystallinity of the synthesized samples was confirmed using X- mv
va
ray diffraction (XRD) analysis (XRD Xpert Pro Panalytical). The
morphology and elemental distribution of the prepared materials were I × Δt
characterized using field emission scanning electron microscopy qc = (2)
m
(FESEM, with a ZEISS Sigma 300 microscope), along with energy-
dispersive X-ray spectroscopy (EDS) elemental maps. High-resolution Csp × ΔV2
E= (3)
transmission electron microscopy (HR-TEM) equipped with an Energy 2 × 3.6
Dispersive X-ray (EDX) system was also employed. X-ray photoelectron
spectroscopy (XPS) analysis was conducted using a Bes Tek Germany E × 3600
P= (4)
instrument to investigate the elemental composition and valence states Δt
of the target elements. The specific surface areas and pore-size distri­ where Cs (C g− 1 for CV tests), and qc (C g− 1 for GCD tests) are the specific
butions were analyzed using the Brunauer-Emmett-Teller (BET, BET capacity, E (W h kg− 1), P (W kg− 1), energy density, and power density,
BELSORP Mini II) method and the Barrett-Joyner-Halenda (BJH) model, respectively. In the above equations, I (A) is the discharge current, Δt is
respectively, through nitrogen sorption measurements conducted at 77 the discharge time, m (g) weight of active material, ϑ scan rate, and ΔV
K. is the operating potential window (V). Under an open-circuit voltage,
electrochemical impedance spectra (EIS) were measured at frequency
2.2.2. Electrochemical measurements ranges of 100 kHz to 0.01 Hz. To evaluate the electrochemical perfor­
For electrochemical evaluation of the target electrode, an electro­ mance of the MCS/NF electrode in practical applications, an asymmetric
chemical workstation (Autolab PGSTAT 302 N Model potentiostat/gal­ supercapacitor (ASC) was assembled. The ASC consisted of MCS/NF as
vanostat (Eco-Chemie, Netherland) that was controlled by NOVA 2.1.1 the positive electrode and activated carbon (AC) as the negative
software) was used and the electrochemical tests (three-electrode and electrode.
two-electrode) were performed in KOH aqueous electrolyte (3 mol L− 1).
The working electrode was NF (1 cm2) covered with the synthesis active

4
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 5. Elemental mapping images for CoMn-S.

3. Results and discussion The co-precipitation reaction of OH− anion with Co2+ , and Mn2+
cations have led to the formation of CoMn-LDH nanoflake structures
3.1. Characterization according to the following reaction on the NF surface [28,29].

Co2+ + Mn2+ + nOH− →CoMn − LDH (6)


Fig. 1 schematically illustrates the synthesis steps of CMS as a posi­
tive electrode. Firstly, the Electrodeposition of Co–Mn LDH is carried Applying a constant potential in the initial phase results in a decrease
out using the Chronoamperometry method by applying a constant po­ in current to its minimum value, indicating the formation of an electric
tential of − 1 V for a duration time of 400 s to allow Mn2+, Co2+, and double-layer charge. This is followed by the electrodeposition of elec­
OH− ions in solution to be deposited onto the NF substrate, resulting in troactive species onto the electrode surface, as depicted in Fig. 2a (zone
Co-Mn− LDH 3D interconnected sheet-like structure. The thickness of a). Subsequently, the current starts to increase, marking the initiation of
the deposited layer is controlled by the applied potential and time to the nucleation reaction, which is the rate-limiting step. During this stage
ensure uniform optimized film formation over the electrode surface. (zone b), both the nucleus growth and the formation of nanoparticles
The NF substrate gradually changes color to green-blue during LDH and nanosheets occur. This process continues until the end of the 400-
deposition, indicating the presence of Co-Mn-OH. Subsequently, the Co- second duration (zone c) [30].
Mn-LDH electrode is exposed to an ethanolic solution of the BTC linker, A three-step mechanism has been proposed for the conversion of
resulting in the formation of a Co-Mn-BTC MOF film on the NF surface CoMn-LDH to CoMn-MOF, involving pseudomorphic transformations,
through an acid-base reaction between LDH and BTC. Finally, the Co- crystal growth, and Ostwald ripening processes [31]. Upon exposure of
Mn-BTC MOF precursor undergoes an anion-exchange reaction in an LDH to the linker, MOF crystals rapidly form on the surface of LDH
electrochemical conversion process, transforming into the CMS struc­ nanosheets, leading to an increase in the thickness of the layers. Initially,
ture while retaining the morphology. the conversion from LDH to MOFs primarily occurs at the interface of the
By applying an optimized fixed potential of − 1 V to the working nanosheets, with a “dissolution-redeposition” mechanism dominating
electrode, OH− ions are produced on the surface of the electrode as a the conversion process. As the reaction time progresses, the MOFs grow
result of NO−3 reduction in accordance with the following reaction [28]. on the nanosheets, resulting in an increase in both the diameter and
roughness of the sheets. At 8 min, vertical nanosheets are no longer
NO−3 + H2 O + 2e− →NO−2 + 2OH− (5)

5
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 6. XPS survey spectra (a), Co 2p pattern (b), Mn 2p pattern (c), S 2p pattern (d) of MnCo-S.

observed, and the surface becomes flat and densely covered with MOFs. (006), (009), and (110) crystal planes are observed, which is consistent
After 9 min, MOF crystals start to grow in the form of rods. This growth with hydrotalcite-like morphology (JCPDS: 10-0144) [33,34]. The XRD
process continues for 20 min, and subsequently, no significant changes pattern of the synthesized MOF crystals indicates 2θ values similar to the
in size and morphology are observed. This stage represents the domi­ values of the major peaks reported in references [35–37] for Co-BTC,
nance of Ostwald ripening in the process. Mn-BTC, and CoMn-BTC (Fig. 3b). The electrochemical conversion sul­
During the electrochemical conversion, MOF (EC-MOF) step, optimal furized product shows diffraction peaks at 34.4◦ , 38◦ , 58.7◦ , and
conditions (scan rate = 0.2 V s− 1, potential window − 1.5–0.2 V, and 60.8◦ corresponding to (311), (400), (440), and (444) crystal planes of
number cycle = 750) are applied using the cyclic voltammetry method. the orthorhombic CMS (JCPDS: 23-1237), respectively (Fig. 3c) [38].
The OH− formed as a result of water oxidation causes the hydrolysis of These results demonstrate the successful synthesis of the CMS structure
Thiourea and then produces S2− . Then, S2− performs an ion exchange on NF.
with the linker and replaces it, and Co-Mn-S is formed. The color of the To study the morphology of product samples such as CoMn-LDH,
electrode turned black once the ion was exchanged, which is a charac­ CoMn-MOF, and CMS synthesized on NF, FE-SEM images of prepared
teristic of Co-Mn− S. Fig. 2b shows the CV curves (initial 250) of the and evaluated (Fig. 4(a–c)). FE-SEM images (Fig. 4a) of CoMn- LDH
formation of CMS due to the electrochemical conversion of CoMn-MOF reveal a hollow structure of uniformly assembled nanosheets formed on
on NF in an aqueous solution containing 1 M KCl (supporting electro­ NF, and the thickness of these nanosheets is approximately less than 50
lyte) and 0.5 M Thiourea (sulfur source) with a scan rate of 0.2 V s− 1 and nm. These interconnected and 3D network structures with high open
saturated with N2 gas. As can be seen in the figure, in addition to the space could help improve the reaction of the linker with Co and Mn for
reduction-oxidation peaks, their shift towards positive and negative creating MOF. Fig. 4b displays the synthesized CoMn-MOF with a typical
potentials with the progress of the reaction indicates the formation of rod-like structure with average thicknesses of about 800 nm × 350 nm.
the electrochemical active substance. Since in this potential range, Higher magnification shows that the synthesized MOF contains smooth
Thiourea does not show any peaks, the presence of these peaks in CV surfaces. After electrochemical conversion and ion exchange, the CMS
indicates the formation of CMS. sample shows a cauliflower-like structure with a high surface area
A set of characterization techniques have been used to confirm the (Fig. 4c). Furthermore, Fig. 5 illustrates the elemental mapping result of
construction of CMS on NF. X-ray diffraction (XRD) investigation was CMS, clearly depicting the signals of Mn, Co, and S elements, along with
used to study the phase information and crystal structure of the syn­ their respective elemental distributions.
thesized products. Fig. 3(a–c) shows the glancing angle X-ray Diffraction The surface chemical composition and elemental valence states of
(GAXRD) pattern of the fabricated electrodes (CoMn-LDH, CoMn-MOF, CMS were characterized using XPS analysis, as shown in Fig. 6a–d. The
and CMS). For all three synthesized electrodes, three peaks with high full survey spectrum (Fig. 6a) (1000–0 Ev) reveals the coexistence of Mn,
intensity at 2θ values of 44.8, 52.3, and 76.7, which attributed to the Co, S, C, and O elements. The O element may be attributed to water
crystal planes of (111), (200), and (220) of the NF substrate [32]. Fig. 3a molecules, while the C element suggests the presence of organic mate­
presents the XRD pattern of the MnCo-LDH precursor, in which the rials absorbed on the sample surface. High-resolution XPS analysis
diffraction peaks at 2θ extent of 11.9, 26.4, 39.2, and 68.6 by the (003), further investigated the elemental composition. Two significant peaks in

6
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 7. (a) TEM image of CoMn-S separated from the NF, (b) HR-TEM images, (c) SAED pattern, and (d) EDX images of the CoMn-S.

the binding energy range of 810 to 760 (eV) indicate the presence of CoMn-LDH, CoMn-MOF, and CoMn-S are around 134.86, 110.71, and
Co2p1/2 and Co2p3/2, and each of these peaks can be indexed at 805, 197 m2 g− 1, respectively. The pore size distributions are given in the
798, 782, and 772 (eV) respectively, which indicate the presence of the inset of figures. The pore size for the samples is mainly centered at 5–20
+3 and +2 oxidation states in CMS (Fig. 6b) [39]. Fig. 6c illustrates how nm coordinated with the property of mesoporous materials. According
the Mn2p spectra (with two shakeup satellites) can be deconvoluted into to the test results, the pronounced enhancement of the SSA for CoMn-S is
two peaks at 644.415 and 626.654 eV for the Mn 2p3/2 and Mn 2p1/2, largely attributed to the complex composition and hierarchical hollow
indicating the coexistence of +2 and +3 valence states of Mn species nanostructure, which is beneficial to provide abundant electrochemical
from the product [39]. The S 2p region’s core-level spectrum is depicted accessible active sites for the intimate contact of electrode/electrolyte
in Fig. 6d, where S 2p1/2 and S 2p3/2 are represented by binding energies and quickening the ions transfer and storage. These results demonstrate
of 165.22 and 163.33 eV, respectively. These results supported the that the as-prepared CoMn-S nano architectures have been successfully
synthesized CMS material and are compatible with the mapping study prepared and can be used for high-performance positive electrodes in
[40]. The detailed intrinsic morphological characteristics and crystal­ hybrid SCs.
linity properties of the CoMn-S were further confirmed by HR-TEM,
SAED, and EDS analysis, which are exhibited in Fig. 7. For the HR- 3.2. Electrochemical measurements
TEM analysis, the sample was prepared by peeling out the CoMn-S
from NF using scratch. The collected sample was then dispersed in 3.2.1. Three electrode tests
DIW under ultrasonication, dropped on a copper grid, and introduced The electrochemical performance of the synthesized electrode for
into the HR-TEM chamber. From the low-magnification TEM image in supercapacitor applications was evaluated using a three-electrode sys­
Fig. 7a, rod-like CoMn-S with hierarchical and intertwined nano- tem. It was observed that the specific capacity of CMS depends on the
architectured morphology were presented. The d-spacing from the lat­ Co: Mn ratio. To investigate this relationship, different Co: Mn ratios
tice fringes of Fig. 7b is about 0.18, 0.23, and 0.36 corresponding to the (3:0, 2:1, 1:1, 1:2, and 0:3) were prepared, and the results of these ex­
(440), (400), and (311), respectively plane of CoMn-S [39,41,42]. The periments are depicted in Fig. 9(a–b). The charge storage capacity is
representative selected area electron diffraction (SAED) pattern of related to the area enclosed by the cyclic voltammetry (CV) curve and
CoMn-S showed ring pattern spots, indicative of the polycrystalline the discharge time obtained from the galvanostatic charge/discharge
nature (Fig. 7c). Moreover, the energy dispersive X-ray (EDX) spectrum (GCD) curve. As shown in Fig. 9a, the CV curve of the CMS sample with a
CoMn-S (Fig. 7d) unambiguously confirms that the structure is Co/Mn ratio of 2:1 exhibits the largest area, indicating the highest
composed of Co, Mn, and S. The BET analysis and pore size distribution charge storage capacity among the tested samples. The specific capacity
were recorded by N2 adsorption-desorption isotherms at 77 K. The re­ initially increases with the introduction of manganese ions to the cobalt
sults are shown in Fig. 8(a–c). We observed that all products possess the sulfide structure, reaching a maximum at the Co/Mn ratio of 2:1, and
typical type-IV curve accompanied by H3 hysteresis characteristic loops then decreasing. This suggests that the incorporation of manganese ions
within a high relative pressure range of 0.5–1.0 P/P0, implying the na­ into the cobalt sulfide structure has a synergistic effect, leading to an
ture of the mesoporous structure. The specific surface areas (SSA) of improvement in the specific capacity. The observed effect can be

7
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 8. N2 adsorption/desorption isotherms for (a) CoMn-LDH, (b) CoMn-MOF, and (c) CoMn-S (the insets show pore size distribution).

attributed to two reasons. Firstly, the combination of CMS with a mixed electrodes exhibit reversible redox signals of the active species in the
composition of cobalt and manganese sulfides results in a richer redox prepared material, indicating a Faradic capacitive behavior character­
reaction compared to monometallic sulfides. This enhances the elec­ istic of a battery-type electrode. The CV curve areas of electrode MCS/
trochemical performance and contributes to the higher specific capacity. NF are significantly larger than those of the other electrodes, indicating
Secondly, the increase in the amount of manganese in Co-sulfide leads to a higher charge storage capacity. This can be attributed to the distinctive
a decrease in the degree of crystallization. This reduced crystallinity can structural design of MCS, which ensures excellent adhesion to the NF
benefit the specific capacity as highly crystalline structures may substrate, as well as the highly porous structure derived from the MOF
encounter limitations in ion penetration and diffusion, hindering the template. These factors create more active sites for redox reactions and
expansion and contraction processes. Therefore, the presence of man­ increase the electrode/electrolyte surface contact. Furthermore, the
ganese helps alleviate diffusion limitations and improves the specific presence of sulfide in the active substance structure enhances the con­
capacity [43,44]. To further explore the capacitance property of as- ductivity and electrochemical efficiency of the prepared electrode. It is
obtained electrodes, the GCDs were tested with a potential window of important to note that the electrochemical capacity of CMS is derived
0–0.4 V. Fig. 9b displays the GCD traces of as-prepared five electrodes at from the quasi-reversible electron transfer processes, which incorporate
5 A g− 1. The CMS electrode with a Co/Mn ratio of 2:1 clearly exhibits the the redox couples Co2+/Co3+ and Mn2+/Mn3+/Mn4+. Therefore, the
longest discharge time compared to the other electrodes at the same charge storage mechanism for the suggested electrode is correlated to
current density, implying the highest value of specific capacitance. The the following reactions [45].
CV curves of CMS/NF at a scan rate of 5 V s− 1 are compared with those of
Mn Co S + OH− + H2 O ↔ MnSOH + CoSOH + 2e− (7)
MnCo MOF/NF and MnCo-LDH/NF, as shown in Fig. 10a. All three

8
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

1
Fig. 9. (a) CV curves, and (b) GCD of CoMn-S electrodes for the five Co: Mn ratios tested (acquired at a sweep rate of 5 mV s− and 5 A g− 1, respectively).

MnSOH + OH− ↔ MnSO + H2 O + e− (8) and inner pore surfaces. This low accessibility of OH− ions resulted in
the decrease of the specific capacitance value towards higher scan rates
CoSOH + OH− ↔ CoSO + H2 O + e− (9) [46].
To investigate the kinetics of charge storage and distinguish between
Consistent with the CV results, a comparison of the galvanostatic capacitive behavior and diffusion-controlled processes, the CV curve
charge/discharge (GCD) tests (Fig. 10b) of the three electrodes dem­ data at different scan rates were analyzed.
onstrates that the proposed CMS electrode has a higher energy storage According to the power law equation (Eq. (10)) [47]:
capacity compared to the other two electrodes.
To further evaluate the electrochemical efficiency, electrochemical i = aϑb (10)
impedance spectroscopy (EIS) measurements were conducted. Nyquist
plots were obtained in a KOH 3 M electrolyte solution within a frequency log(i) = blogϑ + loga (11)
range of 0.01 Hz to 100 kHz at an applied potential of 0.3 V. These The parameters of a, and b are adjustable, where b is the slope, and a
measurements allow the characterization of the intrinsic resistance and is the intercept of the linear plot of log (i) vs. log (ϑ). According to the
electrochemical kinetics of the electrodes. Fig. 10c shows the Nyquist Randles–Sevcik equation (Eq. (12)) [48], when b = 0.5, the current is
plots of CoMn-LDH, CoMn-MOF, and CMS electrodes with the equiva­ proportional to ϑ1/2 :
lent circuit model provided as an inset. The Nyquist plots consist of two
( )1/2
main parts a semicircular region which shows the charge transfer αnf
i = nFAC* D1/2 ϑ1/2 π1/2 X(bt) (12)
resistance (Rct) at high frequencies, and the Warburg line which occurs RT
at low frequencies. The Rct value of CMS (0.52 Ω) is smaller than CoMn-
The diffusion-controlled nature of the current response suggests a
MOF (0.6 Ω), and CoMn-LDH (0.54 Ω), respectively, indicating a faster
faradaic intercalation mechanism, as indicated by Eq. (13). If b = 1, the
charge transfer process and good electrical conductivity. The Warburg
capacitive response can be represented by the following equation, as the
line shows the resistance of the diffusion rate of electrolyte ions during
capacitive current is proportional to the sweep rate [49].
the redox reaction process, which for the CMS sample is lower than the
other two species and indicates the optimal diffusion rate of electrolyte i = ϑCd A (13)
ions and excellent electron transfer kinetics. Upon increasing the scan­
ning rate from 1 to 75 mV s− 1, a series of CV curves with a couple of Fig. 11c presents the data analyzed for anodic and cathodic peaks for
redox peaks was generated (Fig. 11a). It is noteworthy that the anodic the CMS electrode. The b-value for the anodic and cathodic peaks is 0.54
peaks shift towards positive potentials, while the cathodic peaks shift and 0.61 respectively, which indicates the coexistence of two kinds of
towards negative potentials. This behavior is attributed to the difference mechanisms in the electrochemical process of energy storage.
in the diffusion rate of the electrolyte ions, which is not sufficiently fast To quantitatively resolve the contribution from capacitive and
to facilitate the electrochemical processes within the active electrode battery-like processes in the total charge, the following general rela­
material at higher scanning rates. tionship which holds for any material regardless of the charge transfer
The specific capacity, Cs (C g− 1), was calculated using Eq. (1): The process can be used:
Specific capacity of CMS electrode at a scan rate of 2, 5, 8, 10, 15, 25, 50, 1
i (V) = k1 ϑ + k2 ϑ2 (14)
and 75 mV s− 1 was 1110.8C g− 1, 1028.9C g− 1, 948.6C g− 1, 910.9C g− 1,
836.9C g− 1, 743.5C g− 1, 570.1C g− 1, 425.5C g− 1, respectively. Fig. 11b where k1 and k2 denote the coefficients for capacitive and battery-like
shows the specific capacity curve as a function of the scan rates. The currents, respectively. According to Eq. (14), for a redox reaction
specific capacitance values decrease with increasing scan rates. At high 1
limited by semi-infinite linear diffusion, the i(V) varies with ϑ2 , and for a
scan rates, the OH− ions access the outer surface of the electrode
capacitive process, the i(V) varies directly with ϑ. Eq. (14) can be
whereas, at low scan rates the OH− ions intercalate well with the outer

9
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 10. (a) Comparison of CV curves of typical CoMn-LDH, CoMn-MOF, and CoMn-S electrodes at scan rate 5 mV s− 1, (b) Comparison of GCD curves of typical
CoMn-LDH, CoMn-MOF, and CoMn-S electrodes at a current density of 5 A g− 1, (c) EIS spectra for the CoMn-LDH, CoMn-MOF, and CoMn-S electrodes. Insets:
equivalent circuit for the Nyquist plots.

rearranged as: behavior, as depicted in the figure, with two distinct voltage plateaus.
These plateaus correspond to the cyclic voltammetry (CV) plots and
i(V)
(15) indicate the electrochemical performance of the active material as a
1
1 = k1 ϑ2 + k2
ϑ2 battery-type electrode. Battery-type materials and pseudocapacitors can
The parameters k1 and k2 values for each electrode are calculated be distinguished based on their cyclic voltammetry (CV) curves. Battery-
from the slope and y-interception of the plot by linearly fitting the i(V)
1
type materials exhibit faradic redox peaks that are separated by a sig­
1
ϑ2
nificant potential difference, while pseudocapacitor materials display
versus ϑ . The contribution ratio of the surface-controlled and diffusion-
2
completely reversible redox peaks. Battery-type materials often show
dominated process at a series of scan rates from 1 to 75 mV s− 1 are asymmetrical phases in their CV curves, leading to distinct charge-
observed in Fig. 11d. The results show that the capacitive contribution discharge curves with voltage plateaus. In contrast, pseudocapacitors
ratio of CMS electrode gradually increases from 26.7 % to 75.9 % as the exhibit more symmetrical behavior in their CV curves and do not typi­
increasing scan rates, which further indicate the electrochemical process cally display voltage plateaus in their charge-discharge curves [52].
is mainly dominated by diffusion-controlled mechanism [50,51]. Based on the GCD curve at the current density of 1, 2, 4, 8, 15, 20, and 25
In Fig. 12a, the GCD curves of the MCS/NF electrode are shown, A g− 1, the specific capacity was calculated as 1091, 1020, 982, 885, 762,
allowing for the evaluation of the specific capacity at various current 692, and 635C g− 1, respectively. The results of the three electrodes are
densities ranging from 1 to 25 A g− 1. The curves exhibit non-linear shown in the graph of Fig. 12b. The specific capacity decreases as the

10
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 11. (a) CV curves at various scan rates from 1 to 75 mV s− 1, (b) variation in the specific capacity as a function of scan rate, (c) logarithm relationship between
peak current density and scan rate, and (d) the charge storage contribution for CoMn-S at different scan rates.

Fig. 12. (a) GCD profiles at different current densities of CoMn-S electrode, (b) variation of specific capacity with a current density of CoMn-LDH, CoMn-MOF, and
CoMn-S electrodes.

current density increases, which may be related to the corresponding demonstrates the electrode’s ability to maintain its performance and
speed limit of electron transfer and reduced electrolyte ion diffusion retain a significant portion of its capacity under demanding operating
coefficient at high current densities in electrical materials. Indeed, the conditions. The following are the main reasons behind the proposed
CMS/NF electrode has shown excellent behavior and durability even at electrode’s advantage in terms of electrochemical properties: Firstly, the
high current densities. The capacity of the electrode remains at 58.2 % of direct synthesis of the CMS active material on current collectors elimi­
its initial value when subjected to 25 times the current intensity. This nates the need for organic binders and conductive additives. This not

11
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

1
Fig. 13. (a) cycle stability of CoMn-LDH, CoMn-MOF, and CoMn-S electrodes at 25 A g− current density; (b) SEM image of CoMn-S electrode after 7000 cycles.

only reduces dead volume and improves the utilization of active mate­ supercapacitor electrode and showed a specific capacity of 375.2C g− 1 at
rials but also enhances electron transfer from current collectors to active 1 A g− 1 [57].
materials. Secondly, the construction of the MOF-derived CMS results in Furthermore, the unique structure of CMS enables efficient ion and
a three-dimensional porous network. This network facilitates the pene­ electrolyte transfer, minimizing concentration polarization and
tration of the electrolyte (KOH) and increases the electrode/electrolyte enhancing the rate capability of the supercapacitor. The incorporation of
contact surface area. Consequently, more active sites are available, manganese, cobalt, and sulfide elements in the active material structure
leading to an increased capacity for energy storage. One of the unique synergistically improves conductivity, stability, and energy storage ca­
characteristics of supercapacitors is that they are expected to endure a pacity. This synergistic effect enhances the overall efficiency of the
long cycle life. Therefore, the cycling stability of the synthesized CoMn- supercapacitor, making it a favorable choice for energy storage
LDHs, CoMn-MOF, and CoMn-S nanomaterials was tested for 7000 applications.
continuous galvanostatic charge-discharge cycles at 25 A g− 1. As pre­
sented in Fig. 13a, 85 % of the initial capacitance for MnCo-S remained 3.2.2. Two electrode test
after 7000 charge-discharge cycles, which was better than that of the In order to evaluate the performance of the introduced electrode in
other samples, MnCo-MOF (80 %), and MnCo-LDH (75.0 %). the specific practical applications, asymmetric supercapacitor (ASC) devices were
capacity decrease of the CoMn-LDH electrode in the initial cycles can be assembled. The ASC devices consisted of CMS/NF as the positive elec­
attributed to the instability and destruction of the structure [53,54]. In trode, AC/NF as the negative electrode, and a filter paper separator
CoMn-MOF and CoMn-S electrodes, the initial increase is related to the membrane (Fig. 14a). Achieving maximum efficiency in an ASC requires
exposure of the active sites to the electrolyte ions with increasing cycle a balanced charge storage between the positive and negative electrodes
number, and then it becomes stable. SEM images of the MnCo-S after the (q+ = q− ). The mass ratio of the active material on the two electrodes
cycling measurements show that the morphology and structure of was calculated using the following equations, based on the charge bal­
MnCo-S are mostly preserved, even after 7000 cycles (Fig. 13b). These ance equation:
performance parameters are better than some of those reported for MOF-
q = Csp × ΔV × m (16)
derived electrode materials. Xuemin Yin et al. [55] have used metal-
organic frameworks (MOF) by a multi-step method to prepare hierar­ m+ C− × ΔV−
chical Co3O4@NiCoLDH nanosheets (NSs) on carbon cloth. A specific = (17)
m− C+ × ΔV+
capacity of 850C g− 1 was measured at 1 A g− 1. A mesoporous and hollow
Ni-Zn-Co-S nano sword arrays (NSAs) on nickel foam (NF) substrate was In Eq. (16), the amount of charge stored in a single electrode is Csp ,
prepared, which involves the hydrothermal growth of bimetallic Zn-Co- and the operating potential window and the mass of the electrode are Δ
ZIF NSAs on NF and then transforms them into hollow Ni-Zn-Co-S NSAs V and m, respectively. By calculating the capacity of each MCS electrode
through the sulfurization process were demonstrated with a specific and AC electrode for charge balancing in two electrodes, a mass ratio of
capacity of 1.11 mA h cm− 2 at the current density of 10 mA cm− 2 (358.1 about 1:5 was achieved, yielding a final mass value of 6 mg for two
mA h g− 1) by Youzhang Huang et al. [56]. Henan Jia et al. developed electrodes. Fig. 14b represents the CV curves of CMS and AC electrodes
MOFs-derived quadruple-shelled hollow CoS2 dodecahedrons to use a after balancing in different potential windows of − 0.1 to 0.55 and − 1 to

12
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 14. (a) Schematic diagram of the assembled CMS//AC ASC device; (b) CV curves of the AC electrode (− 1 to 0 V) and CMS (− 0.1–0.55 V) in three-electrode cell
configuration at a scan rate of 5 mV s− 1, (c) CV curves of the CMS//AC ASC device in different voltage windows at a scan rate of 5 mV s− 1, (d) CV curves of the CMS//
AC ASC device at different scan rates.

Fig. 15. (a) GCD profiles at different current densities, (b) variation of specific capacity with current density.

13
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

Fig. 16. (a) Ragone plot of MnCo-S//AC device, (b) the cycling stability and corresponding Coulombic efficiency of the ASC device at 15 A g− 1, (the inset shows the
last 16 cycles of the ACS).

Fig. 17. Images of the green LED lighted by two ASC in series. (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)

0 V, respectively, at a scanning rate of 5 mV s− 1. Fig. 14c illustrates a appropriate GCD curves at various current densities ranging from 1 to 15
sequence of CV tests in various potential windows ranging from 0–1 to A g− 1. The GCD curves are almost symmetric, suggesting that the device
0–1.6 V at a scanning rate of 5 mV s− 1 to improve and achieve the is reversible and has a low equivalent series resistance. Based on Eq. (2),
optimal operating potential window of the ACS device. The operational the specific capacity of the ASC device could be evaluated (m is the sum
potential window has been extended to (0–1.6 V), and as there are no of CMS and AC), which are 189.3, 126.3, 94.6, 82, and 75 F g− 1 at 1, 2, 5,
visible cut-off voltages for the positive or negative directions of oxygen 10, and 15 A g− 1, respectively (Fig. 15b). Energy density and power
or hydrogen evolution, this range of potential is suitable for the built-in density are two essential characteristics used to evaluate the practical
device. Furthermore, the combination of the EDLC mechanism in the AC use of energy storage devices and compare them to other suggested
electrode and the battery-type mechanism in the CMS electrode con­ systems. Using Eqs. (3) and (4), when the E is changed from 84.14 to
tributes to the observed CV curves, which exhibit a pair of broad redox 33.3 (W h kg− 1), the P is changed from 1191.4 to 15,000 (W kg− 1).
peaks. This behavior is a result of the unique electrochemical properties Ragone’s plot depicts the findings of this investigation and their
and charge storage mechanisms of each electrode material. The CV tests comparability with previous studies based on ASC devices based on Mn
for the manufactured device at different scanning rates ranging from 1 to and Co chalcogenide (Fig. 16a) [39,58–62]. In addition, the introduced
50 mV s− 1 show that the shape of the CVs remains constant at 50 times device has a higher energy density and power density than some similar
the scanning rate, indicating excellent electrochemical reaction kinetics devices. Cycling stability and Coulombic efficiency are other factors for
and good capacitive behavior, which is an important parameter in en­ practical applications that were investigated in this study using repeated
ergy storage systems (Fig. 14d). Fig. 15a presents the ASC device’s GCD at a current density of 15 A g− 1 (Fig. 16b). As can be observed, after

14
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

7000 cycles, the capacity retention and Coulombic efficiency is around [8] J.S. Lee, D.H. Shin, W. Kim, J. Jang, Highly ordered, polypyrrole-coated Co (OH) 2
architectures for high-performance asymmetric supercapacitors, J. Mater. Chem. A
92 % and 99 % respectively. Finally, by connecting two ASC devices in
4 (2016) 6603–6609.
series, a small green LED was successfully lighted for almost 17 min, as [9] J. Qi, Y. Chang, Y. Sui, Y. He, Q. Meng, F. Wei, et al., Facile synthesis of Ag-
demonstrated by Fig. 17. decorated Ni3S2 nanosheets with 3D bush structure grown on rGO and its
application as positive electrode material in asymmetric supercapacitor, Adv.
Mater. Interfaces 5 (2018) 1700985.
4. Conclusions [10] G. Nagaraju, S.M. Cha, S.C. Sekhar, J.S. Yu, Metallic layered polyester fabric
enabled nickel selenide nanostructures as highly conductive and binderless
electrode with superior energy storage performance, Adv. Energy Mater. 7 (2017)
In summary, we synthesized a binder-free CMS with an effective
1601362.
MOF-derived approach in three steps. An electrosynthesis approach was [11] C. Li, J. Balamurugan, D.C. Nguyen, N.H. Kim, J.H. Lee, Hierarchical
employed to grow Co-Mn-MOF as a porous metal precursor, then a manganese–nickel sulfide nanosheet arrays as an advanced electrode for all-solid-
controlled electrochemical method for converting MOF into the corre­ state asymmetric supercapacitors, ACS Appl. Mater. Interfaces 12 (2020)
21505–21514.
sponding 3D hierarchically porous amorphous CMSx. The Co/Mn ratios [12] M. Moradi, A. Afkhami, T. Madrakian, H.R. Moazami, Hydrothermal synthesis of
and reaction conditions were optimized for the formation of MOF as a nanocages of Mn-Co Prussian blue analogue and charge storage investigation of the
precursor and its conversion to Co-Mn-S (CMS) to achieve a hierarchical derived Mn-Co oxide@/rGO composites, FlatChem 32 (2022), 100350.
[13] J. Xiao, L. Wan, S. Yang, F. Xiao, S. Wang, Design hierarchical electrodes with
architecture with the largest contact surface of active material and highly conductive NiCo2S4 nanotube arrays grown on carbon fiber paper for high-
electrolyte and high electrical conductivity during the synthesis process. performance pseudocapacitors, Nano Lett. 14 (2014) 831–838.
Based on this platform, the CMS was applied and validated as a positive [14] Y. Zhu, X. Ji, Z. Wu, Y. Liu, NiCo2S4 hollow microsphere decorated by acetylene
black for high-performance asymmetric supercapacitor, Electrochim. Acta 186
electrode in SCs, which shows a highly capacitive positive electrode (2015) 562–571.
(1091C g− 1 at 1 A g− 1). Additionally, the assembled ASC device (CMS [15] X.Y. Yu, X.W. Lou, Mixed metal sulfides for electrochemical energy storage and
and AC as anode and cathode respectively), exhibits remarkable energy conversion, Adv. Energy Mater. 8 (2018) 1701592.
[16] B. Ameri, S.S.H. Davarani, H.R. Moazami, H. Darjazi, Cathodic electrosynthesis of
and can exhibit ultrahigh energy density (84 W h kg− 1 at a power
ZnMn2O4/Mn3O4 composite nanostructures for high performance supercapacitor
density of 1191.4 W kg− 1), stable electrochemical stabilities (92 % applications, J. Alloys Compd. 720 (2017) 408–416.
capacitance retention even after 7000 cycles). This study not only sug­ [17] S. Cheng, T. Shi, C. Chen, Y. Zhong, Y. Huang, X. Tao, et al., Construction of porous
CuCo2S4 nanorod arrays via anion exchange for high-performance asymmetric
gests an effective strategy for the rational and controllable design of
supercapacitor, Sci. Rep. 7 (2017) 6681.
porous active materials but also opens a new way to use them for high- [18] Y. Zhang, J. Xu, Y. Zhang, Y. Zheng, X. Hu, Z. Liu, Facile fabrication of flower-like
efficiency electrochemical energy storage. CuCo 2 S 4 on Ni foam for supercapacitor application, J. Mater. Sci. 52 (2017)
9531–9538.
[19] A. Pramanik, S. Maiti, T. Dhawa, M. Sreemany, S. Mahanty, High faradaic charge
CRediT authorship contribution statement storage in ZnCo2S4 film on Ni-foam with a hetero-dimensional microstructure for
hybrid supercapacitor, Mater. Today Energy 9 (2018) 416–427.
Mahdi Moradi: Conceptualization, Methodology, Investigation, [20] Y. Wang, C. Xiang, Y. Zou, F. Xu, L. Sun, J. Zhang, Three-dimensional polypyrrole-
enhanced flower-like ZnCo2S4 nanoclusters used as advanced electrodes for
Writing – original draft. Abbas Afkhami: Supervision, Conceptualiza­ supercapacitors, J. Energy Storage 41 (2021), 102838.
tion, Writing – Review & Editing, Critically revised the manuscript. [21] S.-x. Yan, S.-h. Luo, J. Feng, P.-w. Li, R. Guo, Q. Wang, et al., Rational design of
Tayyebeh Madrakian: Supervision, Writing - review & editing, criti­ flower-like FeCo2S4/reduced graphene oxide films: novel binder-free electrodes
with ultra-high conductivity flexible substrate for high-performance all-solid-state
cally revised the manuscript. Hamid Reza Moazami: Supervision, pseudocapacitor, Chem. Eng. J. 381 (2020), 122695.
Writing – Review & Editing, critically revised the manuscript. [22] S. Tang, B. Zhu, X. Shi, J. Wu, X. Meng, General controlled sulfidation toward
achieving novel nanosheet-built porous square-FeCo2S4-tube arrays for high-
performance asymmetric all-solid-state pseudocapacitors, Adv. Energy Mater. 7
Declaration of competing interest (2017) 1601985.
[23] S. Shinde, G. Ghodake, N. Maile, H. Yadav, A. Jagadale, M. Jalak, et al., Designing
of nanoflakes anchored nanotubes-like MnCo2S4/halloysite composites for
The authors declare that they have no known competing financial advanced battery like supercapacitor application, Electrochim. Acta 341 (2020),
interests or personal relationships that could have appeared to influence 135973.
the work reported in this paper. [24] K. Yu, W.M. Tang, J. Dai, Double-layer MnCo2S4@ Ni-Co-S core/shell
nanostructure on nickel foam for high-performance supercapacitor, Phys. Status
Solidi 215 (2018) 1800147.
Data availability [25] V. Gajraj, R. Azmi, S. Indris, C. Mariappan, Boosting the multifunctional properties
of MnCo2O4-MnCo2S4 heterostructure for portable all-solid-state symmetric
supercapacitor, methanol oxidation and hydrogen evolution reaction,
No data was used for the research described in the article.
ChemistrySelect 6 (2021) 11466–11481.
[26] J. Zhao, Y. Wang, Y. Qian, H. Jin, X. Tang, Z. Huang, et al., Hierarchical design of
References cross-linked NiCo2S4 nanowires bridged NiCo-hydrocarbonate polyhedrons for
high-performance asymmetric supercapacitor, Adv. Funct. Mater. 2210238 (2022).
[27] H. Wang, K. Zhang, Y. Song, J. Qiu, J. Wu, L. Yan, MnCo2S4 nanoparticles
[1] S. Chu, A. Majumdar, Opportunities and challenges for a sustainable energy future,
anchored to N-and S-codoped 3D graphene as a prominent electrode for
Nature 488 (2012) 294–303.
asymmetric supercapacitors, Carbon 146 (2019) 420–429.
[2] M.S. Javed, H. Lei, J. Li, Z. Wang, W. Mai, Construction of highly dispersed
[28] Y. Tang, H. Shen, J. Cheng, Z. Liang, C. Qu, H. Tabassum, et al., Fabrication of
mesoporous bimetallic-sulfide nanoparticles locked in N-doped graphitic carbon
oxygen-vacancy abundant NiMn-layered double hydroxides for ultrahigh capacity
nanosheets for high energy density hybrid flexible pseudocapacitors, J. Mater.
supercapacitors, Adv. Funct. Mater. 30 (2020) 1908223.
Chem. A 7 (2019) 17435–17445.
[29] F. Grote, Z.Y. Yu, J.L. Wang, S.H. Yu, Y. Lei, Self-stacked reduced graphene oxide
[3] G. Liu, C. Kang, J. Fang, L. Fu, H. Zhou, Q. Liu, MnO2 nanosheet-coated Co3O4
nanosheets coated with cobalt–nickel hydroxide by one-step electrochemical
complex for 1.4 V extra-high voltage supercapacitors electrode material, J. Power
deposition toward flexible electrochromic supercapacitors, Small 11 (2015)
Sources 431 (2019) 48–54.
4666–4672.
[4] C. Kang, J. Fang, X. Liu, S. Li, S. Wan, L. Fu, et al., A novel fabricated conductive
[30] Y. Wang, Z. Yin, G. Yan, Z. Wang, X. Li, H. Guo, et al., New insight into the
substrate for enhancing the mass loading of NiCoLDH nanosheets for high areal
electrodeposition of NiCo layered double hydroxide and its capacitive evaluation,
specific capacity in hybrid supercapacitors, Electrochim. Acta 368 (2021), 137621.
Electrochim. Acta 336 (2020), 135734.
[5] G.K. Veerasubramani, K. Krishnamoorthy, S.J. Kim, Improved electrochemical
[31] K. Okada, R. Ricco, Y. Tokudome, M.J. Styles, A.J. Hill, M. Takahashi, et al.,
performances of binder-free CoMoO4 nanoplate arrays@ Ni foam electrode using
Copper conversion into Cu (OH) 2 nanotubes for positioning Cu3 (BTC) 2 MOF
redox additive electrolyte, J. Power Sources 306 (2016) 378–386.
crystals: controlling the growth on flat plates, 3D architectures, and as patterns,
[6] G. Nagaraju, S.C. Sekhar, G.S.R. Raju, L.K. Bharat, J.S. Yu, Designed construction of
Adv. Funct. Mater. 24 (2014) 1969–1977.
yolk–shell structured trimanganese tetraoxide nanospheres via polar solvent-
[32] G.S.R. Raju, E. Pavitra, G. Nagaraju, S.C. Sekhar, S.M. Ghoreishian, C.H. Kwak, et
assisted etching and biomass-derived activated porous carbon materials for high-
al., Rational design of forest-like nickel sulfide hierarchical architectures with
performance asymmetric supercapacitors, J. Mater. Chem. A 5 (2017)
ultrahigh areal capacity as a binder-free cathode material for hybrid
15808–15821.
supercapacitors, J. Mater. Chem. A 6 (2018) 13178–13190.
[7] Z. Supardi, A. Najihah, S. Priyono, B. Prihandoko, Electrochemical performances of
lithium-ion coin cell based on Li4Ti5O12 anode, in: IOP Conference Series:
Materials Science and Engineering, IOP Publishing, 2021, p. 012007.

15
M. Moradi et al. Journal of Energy Storage 71 (2023) 108177

[33] S. Liu, S.C. Lee, U. Patil, I. Shackery, S. Kang, K. Zhang, et al., Hierarchical MnCo- [47] F. Ma, X. Dai, J. Jin, N. Tie, Y. Dai, Hierarchical core-shell hollow CoMoS4@
layered double hydroxides@ Ni (OH) 2 core–shell heterostructures as advanced Ni–Co–S nanotubes hybrid arrays as advanced electrode material for
electrodes for supercapacitors, J. Mater. Chem. A 5 (2017) 1043–1049. supercapacitors, Electrochim. Acta 331 (2020), 135459.
[34] A. Emin, X. Song, Y. Du, Y. Chen, M. Yang, S. Zou, et al., One-step electrodeposited [48] A.J. Bard, L. Faulkner, Electrochemical Methods: Fundamentals and Applications,
Co and Mn layered double hydroxides on Ni foam for high-performance aqueous 409 vol. 410, John Wiley and Sons, N Y, 1980.
asymmetric supercapacitors, J. Energy Storage 50 (2022), 104667. [49] H. Lindström, S. Södergren, A. Solbrand, H. Rensmo, J. Hjelm, A. Hagfeldt, et al., Li
[35] X. Zhang, A. Yuan, X. Mao, Q. Chen, Y. Huang, Engineered Mn/Co oxides + ion insertion in TiO2 (anatase). 2. Voltammetry on nanoporous films, J. Phys.
nanocomposites by cobalt doping of Mn-BTC-new oxidase mimetic for colorimetric Chem. B 101 (1997) 7717–7722.
sensing of acid phosphatase, Sensors Actuators B Chem. 299 (2019), 126928. [50] F. Tu, Y. Han, Y. Du, X. Ge, W. Weng, X. Zhou, et al., Hierarchical nanospheres
[36] F. Zheng, G. Xia, Y. Yang, Q. Chen, MOF-derived ultrafine MnO nanocrystals constructed by ultrathin MoS2 nanosheets braced on nitrogen-doped carbon
embedded in a porous carbon matrix as high-performance anodes for lithium-ion polyhedra for efficient lithium and sodium storage, ACS Appl. Mater. Interfaces 11
batteries, Nanoscale 7 (2015) 9637–9645. (2018) 2112–2119.
[37] N. Yang, H. Song, X. Wan, X. Fan, Y. Su, Y. Lv, A metal (Co)–organic framework- [51] V. Augustyn, P. Simon, B. Dunn, Pseudocapacitive oxide materials for high-rate
based chemiluminescence system for selective detection of L-cysteine, Analyst 140 electrochemical energy storage, Energy Environ. Sci. 7 (2014) 1597–1614.
(2015) 2656–2663. [52] T.S. Mathis, N. Kurra, X. Wang, D. Pinto, P. Simon, Y. Gogotsi, Energy storage data
[38] S.-Y. Hsu, F.-H. Hsu, J.-L. Chen, Y.-S. Cheng, J.-M. Chen, K.-T. Lu, The reporting in perspective—guidelines for interpreting the performance of
supercapacitor electrode properties and energy storage mechanism of binary electrochemical energy storage systems, Adv. Energy Mater. 9 (2019) 1902007.
transition metal sulfide MnCo 2 S 4 compared with oxide MnCo 2 O 4 studied using [53] G. Hu, C. Tang, C. Li, H. Li, Y. Wang, H. Gong, The sol-gel-derived nickel-cobalt
in situ quick X-ray absorption spectroscopy, Mater. Chem. Front. 5 (2021) oxides with high supercapacitor performances, J. Electrochem. Soc. 158 (2011)
4937–4949. A695.
[39] A.M. Elshahawy, X. Li, H. Zhang, Y. Hu, K.H. Ho, C. Guan, et al., Controllable [54] X. Wang, C. Yan, A. Sumboja, P.S. Lee, High performance porous nickel cobalt
MnCo 2 S 4 nanostructures for high performance hybrid supercapacitors, J. Mater. oxide nanowires for asymmetric supercapacitor, Nano Energy 3 (2014) 119–126.
Chem. A 5 (2017) 7494–7506. [55] Y. Xuemin, L. Hejun, Y. Ruimei, L. Jinhua, NiCoLDH nanosheets grown on MOF-
[40] S.Y. Khoo, J. Miao, H.B. Yang, Z. He, K.C. Leong, B. Liu, et al., One-step derived Co3O4 triangle nanosheet arrays for high-performance supercapacitor,
hydrothermal tailoring of NiCo2S4 nanostructures on conducting oxide substrates J. Mater. Sci. Technol. 62 (2021) 60–69.
as an efficient counter electrode in dye-sensitized solar cells, Adv. Mater. Interfaces [56] Y. Huang, L. Quan, T. Liu, Q. Chen, D. Cai, H. Zhan, Construction of MOF-derived
2 (2015) 1500384. hollow Ni–Zn–Co–S nanosword arrays as binder-free electrodes for asymmetric
[41] M. Wang, Y. Lai, J. Fang, F. Qin, Z. Zhang, J. Li, et al., Hydrangea-like NiCo 2 S 4 supercapacitors with high energy density, Nanoscale 10 (2018) 14171–14181.
hollow microspheres as an advanced bifunctional electrocatalyst for aqueous [57] H. Jia, Z. Wang, X. Zheng, Y. Cai, J. Lin, H. Liang, et al., Controlled synthesis of
metal/air batteries, Catal. Sci. Technol. 6 (2016) 434–437. MOF-derived quadruple-shelled CoS2 hollow dodecahedrons as enhanced
[42] S. Peng, L. Li, C. Li, H. Tan, R. Cai, H. Yu, et al., In situ growth of NiCo 2 S 4 electrodes for supercapacitors, Electrochim. Acta 312 (2019) 54–61.
nanosheets on graphene for high-performance supercapacitors, Chem. Commun. [58] J. Xu, Y. Sun, M. Lu, L. Wang, J. Zhang, E. Tao, et al., Fabrication of the porous
49 (2013) 10178–10180. MnCo2O4 nanorod arrays on Ni foam as an advanced electrode for asymmetric
[43] M. Huang, F. Li, F. Dong, Y.X. Zhang, L.L. Zhang, MnO 2-based nanostructures for supercapacitors, Acta Mater. 152 (2018) 162–174.
high-performance supercapacitors, J. Mater. Chem. A 3 (2015) 21380–21423. [59] H. Jia, Y. Song, J. Wu, W. Fu, J. Zhao, X. Liu, A novel P-doped MnCo2S4
[44] G. Zhu, C. Xi, M. Shen, C. Bao, J. Zhu, Nanosheet-based hierarchical Ni2 (CO3) nanoneedles assembled dandelion-like structure for high performance hybrid
(OH) 2 microspheres with weak crystallinity for high-performance supercapacitor, supercapacitors, Mater. Lett. 233 (2018) 55–58.
ACS Appl. Mater. Interfaces 6 (2014) 17208–17214. [60] K. Zarean Mousaabadi, A.A. Ensafi, B. Rezaei, Co3O4/MoCo/layered double
[45] S. Liu, S.C. Jun, Hierarchical manganese cobalt sulfide core–shell nanostructures hydroxide nanosheets for asymmetric supercapacitor, ACS Appl. Nano Mater. 5
for high-performance asymmetric supercapacitors, J. Power Sources 342 (2017) (2022) 8097–8104.
629–637. [61] X. Zhai, H. Pan, F. Wang, X. Gao, Z. Xiong, L. Li, et al., Controlled growth of 3D
[46] S. Vijayakumar, S. Nagamuthu, G. Muralidharan, Porous NiO/C nanocomposites as interpenetrated networks by NiCo2O4 and graphdiyne for high-performance
electrode material for electrochemical supercapacitors, ACS Sustain. Chem. Eng. 1 supercapacitor, ACS Appl. Mater. Interfaces 14 (2022) 18283–18292.
(2013) 1110–1118. [62] Q. Fang, M. Sun, X. Ren, Y. Sun, Y. Yan, Z. Gan, et al., MnCo2O4/Ni3S4
nanocomposite for hybrid supercapacitor with superior energy density and long-
term cycling stability, J. Colloid Interface Sci. 611 (2022) 503–512.

16

You might also like