Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

A Journal of

Accepted Article

Title: Rational designing of porous structured nickel manganese


sulfides hexagonal sheets-in-cage structures as an advanced
electrode material for high-performance electrochemical
capacitors

Authors: Diab Khalafallah, Zongxiao Wu, Mingjia Zhi, and Zhanglian


Hong

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.

To be cited as: Chem. Eur. J. 10.1002/chem.201904991

Link to VoR: http://dx.doi.org/10.1002/chem.201904991

Supported by
Chemistry - A European Journal 10.1002/chem.201904991

Rational designing of porous structured nickel manganese sulfides


hexagonal sheets-in-cage structures as an advanced electrode
material for high-performance electrochemical capacitors

Dr. Diab khalafallah1,2,*, Zongxiao Wu1, Dr. Mingjia Zhi1,*, Prof. Zhanglian
Hong1,*

Accepted Manuscript
1
State Key Laboratory of Silicon Material, School of Materials Science and Engineering, Zhejiang
University, 38 Zheda Road, Hangzhou, 310027, China
2
Mechanical Design and Materials Department, Faculty of Energy Engineering, Aswan University, P.O.
Box 81521, Aswan, Egypt

Email: mingjia_zhi@zju.edu.cn; hong_zhanglian@zju.edu.cn, diab@zju.edu.cn

Abstract
The design of hierarchical electrode comprising multiple components with a high electrical

conductivity and a large specific surface area has been recognized as a feasible strategy to

remarkably boost pseudocapacitors. Herein, we delineate hexagonal sheets-in-cage shaped

nickel-manganese sulfides (Ni-Mn-S) with nanosized open spaces for supercapacitors

application to realize faster redox reactions and a lower charge transfer resistance with a

markedly enhanced specific capacitance. The hybrid was facilely prepared through a two-step

hydrothermal method. Benefiting from the synergistic effect between Ni and Mn active sites with

the improvement of both ionic and electric conductivity, the resulting Ni-Mn-S hybrid displays a

high-specific capacitance of 1664 F g−1 at a current density of 1 A g−1 and a capacitance of 785 F

g−1 is maintained at a current density of 50 A g−1, revealing an outstanding capacity and rate

performance. The asymmetric supercapacitor device assembled with the Ni-Mn-S hexagonal

sheets-in-cage as the positive electrode delivers a maximum energy density of 40.4 Wh kg−1 at a

power density of 750 W kg-1. Impressively, the cycling retention of the as-fabricated device after

10,000 cycles at a current density of 10 A g−1 reaches 85.5%. Thus, the hybrid Ni-Mn-S

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

hexagonal sheets-in-cage with superior capacitive performances holds a great potential as an

effective charge storage material.

Keywords: Nickel-manganese sulfides, hexagonal sheets-in-cage, porous heterostructures,

synergistic effect, electrochemical behavior, asymmetric supercapacitors

Accepted Manuscript
Introduction

Clean energy storage and conversion systems such as fuel cells, solar cells, batteries,

supercapacitors, and etc. have been considered as the frontier of sustainable energy research

owing to the foreseeable energy problems associated with the depletion of conventional fossil

fuels and aggravated environmental pollution [1,2]. Moreover, the increasing demand for such

portable electronic instruments to fulfill the current and future requirements of energy has

motivated a vigorous research interest in recent years due to light weight, zero emission, and

relatively high-power density of these renewable technologies. Among these energy conversion

and storage devices, supercapacitors have been regarded as a privileged technology for energy

storage, but their low power densities and inferior long-term stabilities prohibit their large

commercialized adoption [3,4]. Supercapacitors can be distinguished into two main types

according to the mechanism of charge storage; pseudocapacitors and electrochemical double

layer capacitors (EDLCs). In this context, the charge transfer within the EDLCs takes place via

ion adsorption at electrode/electrolyte interfaces, while in the pseudocapacitors the charge is

stored due to rapid and reversible redox reactions onto electrode surface [5]. Transition metals

and conducting polymers-based electrodes are more favourable for pseudocapacitors, whereas,

carbon counterparts like graphene, porous carbon, and carbon nanotubes are examples of

electrode materials for EDLCs, which possess low specific capacitances compared to

pseudocapacitors [6-8]. More importantly, electrode material is a key component to

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

supercapacitors. With the aim of accomplishing a high performance, an intensive research

progress has been afforded throughout fabricating micro/nanostructured electrodes with various

compositions and surface morphologies, however current supercapacitors still need significant

enhancements to further increase their energy densities without infringing their inherent features

[9,10]. Over the past decade, nanomaterials have been commonly investigated as supercapacitor

Accepted Manuscript
electrodes because of their large specific surface area and fast electron/ion diffusion as well.

Recently, transition metals sulfides-based materials have drawn a great importance as promising

alternatives with a reasonable specific capacity and energy density originated from their unique

physical and chemical properties. Recently, metal sulfides containing transition elements

particularly, nickel, cobalt, manganese, and copper showed improved electrochemical properties

with good specific capacitances and stabilities, which are affiliated to their abundant

electrochemistry and rapid redox reactions. Sulfides possess an excellent electrochemical activity

and ionic conductivity due to the substitution of oxygen with sulfur species and large sizes of S 2‒

ions [11]. Binary or ternary transition metal sulfides are more preferable than unitary metal

sulfides and their energy storage performance outperforms the corresponding binary or ternary

transition metal oxides due to improved electronic structure and copious electroactive sites [12-

16]. Moreover, the low electric conductively of monometal sulfides can largely limited their

widespread applications in supercapacitors [17,18], thus it is important to explore new electrode

materials with desirable properties. The coupling of two or more metallic species to render

compounds of multiple components can provide much enhanced specific capacitances at higher

current densities compared to single component sulfide. The synergistic improvements in

properties including chemical performance and stability as well as poisoning resistance are much

higher than those realized by individual components [19]. Furthermore, the composition control

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

of multicomponent sulfides can sufficiently endow a new degree of freedom to promote the

chemical activity through optimizing the electrical conductivity.

Numerous reports tend to combine various sulfides in different compositions to enhance the

charge storage properties of supercapacitors. For example, Shen et al. presented nickel cobalt

sulfide hollow spheres as an electrode material with a high specific capacitance of 1036 F g−1 at a

Accepted Manuscript
current density of 1 A g−1 [20]. Xia et al. fabricated urchin shaped NiCo2S4 nanostructures via

two-step hydrothermal approach. The as-prepared hybrid electrode exhibited a high capacitance

of 1149 F g−1 at 0.5 A g−1 under three electrode configurations [21]. Similarly, Wang et al.

deposited NiCo2S4 nanotube arrays directly onto a carbon fiber paper by hydrothermal treatment

with an excellent capacitance of 0.87 F cm−2 at 4 mA cm−2 [22]. Sahoo and coworkers reported

nanosheets-like nickel cobalt manganese mixed sulfides (NCMS) as a free-standing electrode by

a cathodic electrodeposition route. The self-supported NCMS nanosheets achieved a distinct

specific capacitance of 2717 F g−1 at 1 A g−1 [15]. In another work, Peng et al. synthesized

cobalt–manganese sulfides nanoneedle arrays onto a microporous current collector by multi-step

hydrothermal method. The binder free electrode revealed a high capacity of 0.53 mA h cm−2 at 2

mA cm−2 with a good rate performance after 1500 cycles [23]. By taking advantage of tunable

composition and promoted carrier mobility, binary and trimetallic sulfides often give rise to

intriguing electronic/electrochemical properties. Hence, the exploitation of multi-sulfide

compounds through a delicate integration of capacitive and faradic components conforms to the

prerequisites of sustainable energy storage systems.

We here report the development of binary Ni-Mn-S hexagonal sheets-in-cage by a two-step

hydrothermal route as an advanced charge storage material with expect to improve the active

sites density, electron transfer kinetics, and as well as ion diffusion capability. The collaborative

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

contribution from both components guided the hierarchically porous Ni-Mn-S electrode to

achieve high specific capacitances of 1664 F g−1 at 1 A g−1 and 785 F g−1 at 50 F g−1 with a good

cyclic stability (90.7% capacity retention after 3000 cycles) due to the substantial enhancement

in specific surface area, electron transport, and electroactive sites for redox reactions. Moreover,

the hexagonal nanosheets could prevent the cages from breakage during redox reactions,

Accepted Manuscript
therefore a high utilization of the active components might attained. An asymmetric

supercapacitor investigated the Ni-Mn-S hexagonal sheets-in-cage as cathode material and

showed a high energy at a high power density, proving its promising potential for next-

generation energy storage systems.

Results and discussion

Formation of hierarchical Ni-Mn-S nanostructure

The ion exchange reaction has been considered as an effective pathway to transform

nano/microarchitectures through ecofriendly synthesis approaches [24]. This technique avoids

the usage of toxic sulfur precursors and high temperatures, while maintaining a high

reproducibility and affording scalable production rates. Therefore, this mechanism holds a great

significant in developing novel materials with favourable features by utilizing the purposeful

insertion of corresponding ions [Table S1]. Even the transformation of a single crystal-to-single

crystal may be possible through this strategy [25a].

To the best of our knowledge, this method has not been applied for the preparation of hexagonal

sheets-in-cage-like Ni-Mn-S composite with a unique porous texture as an efficient

supercapacitor electrode. The presented synthesis technique is an ecologically friendly protocol

for sulfurization. The interior components were etched out and as a result, Ni-Mn-S with novel

nano/microarchitectures circumvents the dead volume of the reactive matrix. The fabrication of

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

hierarchically porous Ni-Mn-S hexagonal sheets-in-cage composite was achieved via a facile

hydrothermal treatment of metallic salts and a subsequent anion exchange (e.g. sulfidation

process) (Scheme 1). The hydrothermal hydrolysis of Ni2+and Mn2+ cations in an alkaline

environment can lead to extended chains due to coordination with H+ ions and subsequently

interlinked to generate two dimensional (2D) hexagonal nanosheets. The effect of urea could be

Accepted Manuscript
understood as a self-template or a structure directing agent to induce the growth of Ni-Mn-

precursor through the in-situ reaction between Ni2+/Mn2+ and hydrolysis products (CO32─ and

OH─ anions) [25b,26]. Meanwhile, the NH4F anions play a role in two ways: (i) as another kind

of structure directing agent for the assembly 2D hexagonal sheets into a three-dimensional (3D)

cage structure by providing more active centers for nucleation and growth of mesocrystals during

the hydrothermal reaction; (ii) as a complexing agent to link the Ni2+ and Mn2+ cations and hence

facilitates the formation of bimetallic composition for self-seed establishment [27]. The Ni and

Mn precursor deliberated by urea and NH4F proceeds a localized dissolution-precipitation

processes of Ni2+ and Mn2+ to form Ni-Mn-(OH)x(SO4)y (Fig.S1) [28].

Scheme 1 Fabrication process for 3D heterostructured Ni-Mn-S hexagonal sheets-in-cage by

two-step hydrothermal approach.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Because of Ostwald ripening reaction, the newly generated nuclei tended to integrate with the

neighboring nuclei in order to reduce the interfacial energy, leading to an oriented agglomeration

of saturated nuclei. At last, stable hexagonal sheets were constructed due to fast crystal growth.

Further interplay of components makes the self-assembling behaviour and re-growth processes

happen till the formation of 3D cage superstructures with hexagonal sheets as the building

Accepted Manuscript
blocks. In the present reaction system, it can be suggested that the formation of OH─ is

predominant at low content of urea within the reaction, leading to the existence of Ni-Mn-

(OH)x(SO4)y. The generation of HCO3 might be prohibited because of insufficient CO2 and

CO32─ species [28]. Thereafter, a subsequent hydrothermal sulfidation process was conducted to

convert the obtained Ni-Mn-(OH)x(SO4)y precursor into Ni-Mn-S through etching the interior

components with Na2S while retaining their intrinsic structure. In such reaction, the H2S and HS‒

species were produced as a result of Na2S dissociation inside the solvent into S2‒ and these

species act as a sulfur source during the vulcanization reaction [29]. Interestingly, the interior

mesocrystals and created void spaces provide much more active sites for ease diffusion reactions

between the electrochemically active Ni-Mn-S and electrolyte ions, which represents key points

for better charge storage performance.

Fig. 1A shows the XRD patterns of the as-synthesized products. The observed diffraction peaks

at about 21.6o, 31.7o, 50.1o, and 55.2o can be identified as (101), (110), (113), and (122) planes,

respectively of Ni3S2 (JCPDS No. 44-1418) [30]. Besides, the peaks located at 30.2o, 34.6o,

45.8o, 53.5o, 61o, 65.4o, and 73.1o are belonging to [100], [101], [102], [110], [103], [201], and

[202] planes, respectively of hexagonal NiS phase (PDF#75-0613) [31]. Moreover, the

characteristics existed at ~25.8o, 27.6o, 31.7o, 38.2o, 49.9o, and 55.2o agree well with the [100],

[002], [111], [102], [103], and [201] planes, respectively of γ-MnS phase (JCPDS no. 65-3413)

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

[32]. The absence of other peaks suggests the high purity of the prepared materials. Thus, it is

reasonable to conclude that the Ni-Mn-S composite can fabricated using our developed synthesis

route.

Accepted Manuscript
Fig. 1 (A) XRD patterns of the as-fabricated Ni-Mn-S composite (a) and pristine Ni-S (b). High

resolution spectra of Mn 2p region (B), Ni 2p (C), and S 2p (D) of Ni-Mn-S composite.

The chemical composition of the hybrid Ni-Mn-S hexagonal sheets-in-cage was further verified

by XPS analyses. The full survey scan of the hybrid shown in Fig. S1 displays the existence of

Ni, Mn, S, O, and C elements. The observed C and oxygen signals are mainly scribed to the

reference carbon and partial surface oxidation, respectively. However, the near surface of the

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

hybrid comprises Ni, Mn, and S elements. The Gaussian fitting of the Mn 2p emission spectra

shows two major spin-orbit doublets located at binding energies of 641.6 eV and 653.5 eV

corresponding to Mn 2p3/2 and Mn 2p1/2 of Mn2+, respectively [33,34]. The weak satellite peak

(Sat.) appeared at 646.6 eV is associated to divalent Mn species such as MnO. The

deconvolution of complex Ni 2p features two main spin–orbit lines with two shakeup satellites at

Accepted Manuscript
855.7 eV and 873.7 eV assigned to Ni 2p3/2 and Ni 2p1/2 bands Ni2+, respectively [35]. In the S 2p

spectral region, two main peaks are found at 160.8 eV and 162 eV attributed to the coordination

of sulfur with metal ions and identified as S 2p3/2 and S 2p1/2, respectively. Meanwhile the

characteristic peak at 168.8 eV represents the metal sulfate originated from the surface S atoms

oxidation. Thus, the near surface composition of the hybrid Ni-Mn-S hexagonal sheets-in-cage is

Ni2+, Mn2+, and S2-, which matches well with the XRD results.

The surface configuration, shape, and porous nature of electroactive material are critical issues

influencing its charge storage capability. The SEM image of monometallic Ni-S is illustrated in

Fig. S3. As shown the sulfurized product exhibits 2D hexagonal plate structure. The panoramic

SEM micrographs of Ni-Mn-(OH)x(SO4)y precursor (Figs. 2A-C&S4A-B) and Ni-Mn-S final

product (Figs. 2D-F&S4C-D) demonstrate that both samples show 2D hexagonal sheets

assembled into 3D microcage architectures. The hexagonal sheets are firmly attached and

hosting numerous void spaces, forming a highly open heterostructures, which investigated the

advantages of metallic ions hydrolytic characteristics and morphology-directing agent.

These results indicate that hexagonal sheets could be easily produced in presence of urea and

integrate into giant microcages due to inherent crystal tendency [36]. Such reaction conditions

created an appropriate interlayer H-bonding and subsequently hexagonal sheets were bundled

into a unique hutch-like structure due to possible interfacial tension [37]. However, the most

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

influential parameter on the final morphology is the change in growth rates of the corresponding

crystal faces as the sulfate ions favors to incorporate into the Ni-Mn-(OH)x(SO4)y interlayers

[38,39]. In fact, the growth of hexagonal sheets was controlled by the gradual consumption of

reactants concentrations within the solution, and this can explain the formation of incomplete

crystals at the surfaces of fully-grown building blocks and final formation of cage structures. The

Accepted Manuscript
existence of enough spaces among the adjoining sheets within the hierarchical heterostructures

verifies the inhibition of agglomeration phenomenon. The sheets still remain their homogeneous

hexagonal structure after anion-exchange reactions with Na2S as presented in Figs. 2D-F&S4C-

D. However, the surface morphology of the final product changed slightly. The magnified SEM

images of the Ni-Mn-S hexagonal sheets-in-cage reveal the formation of randomly distributed

nanoparticles and anchored onto the top surfaces and vacant spaces between the sheets

(Fig.2F&S4D). These nanoparticle subunits are attributed to secondary growth and nucleation at

the surface of existing sheets [40a]. The SEM observation illustrates that the hexagonal sheets

were thermally stable and revealed no noticeable morphology collapse (Figs. 2F&S4D). Fig.

2C&F shows the top-view SEM image of sheet-like surface and edges of precursor and final

product with a sheet thickness ranging from 80 to 130 nm. Such flexible sheet-in-cage-like

morphologies with open windows not only provide large contact regions with the electrolyte

during electrochemical redox reactions but also shorten the diffusion pathways of ions, and

hence facilitating the electrochemical reactions within the overall heterostructured material.

Apparently, the formation of a large range of tiny particles with hollow voids during the

vulcanization reaction (as indicated by yellow arrows-Fig. 2F) ensure more reactive sites

participating in redox reactions, guaranteeing a high specific capacity and a good cycling

stability. Most important, the hexagonal sheet matrices-like building blocks might serve as a

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

robust mechanical support to stabilize the entire particles, favoring the superior preservation of

capacitance, particularly at higher current densities. In general, such flexible hexagonal sheets-

in-cage frameworks can adequately enhance the cycling stability of the electrode over an

extended range of continuous charge/discharge processes [25c,40b].

Fig. 2 SEM images of the as-prepared materials with different magnifications. (A-C) Ni-Mn-
Accepted Manuscript
(OH)x(SO4)y precursor and (D-F) Ni-Mn-S final product. The outcomes confirm the formation of

hexagonal sheets building blocks assembled into 3D cage structures (yellow lines in panels-

B&D-F). The surface roughness of the sheets increases after vulcanization reaction.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Detailed micromorphology information of precursor and final product Ni-Mn-S hexagonal

sheets-in-cage frameworks are obtained by TEM. The observations demonstrate the high

flexibility, transparency, and as well as porosity features of the constructed building blocks

which coincide well with the SEM analysis (Fig. 3A-F). For both composites, we can see that the

pronounced sheets internally possess a well-defined hexagonal structure and are robustly

Accepted Manuscript
interconnected by electrostatic interactions such as hydrogen bonding, forming a porous 3D

hexagonal sheets-in-cage assemblies. The porous characteristics could be accomplished by a

mechanism analogous to the Kirkendall reaction, in which the vacancies are generated as a result

of unbalanced counter diffusion inside the reaction interface [41]. In the HR-TEM image of

precursor presented in Fig. 3C, the well-resolved lattice fringes separated by distances of 0.445

nm and 0.385 nm are attributed to [100] and [010] lattice parameters of β-Ni(OH)2, respectively.

As it can be seen from Fig. 3D-E, the vulcanization reaction has no obvious effect upon the

shape of the grown sheets, but the in-situ anion exchange reactions direct the crystallization

towards larger particles. It is reasonably that the S2‒ ions’ reaction at the surface improved the

porous features of the Ni-Mn-S composite due to relatively rapid anion-exchange reactions at the

surfaces.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Accepted Manuscript
Fig. 3 (A-C) Low and HR-TEM images of Ni-Mn-(OH)x(SO4)y precursor. The sample shows a

3D cage-like morphology assembled from a group of thin hexagonal plates/sheets as indicated by

the yellow lines (panel-A). The HR-TEM images presents the resolved lattice fringes as

illustrated by the yellow parallel lines. (D-E) Low resolution-TEM analysis and (F) HR-TEM

image of the composite Ni-Mn-S hexagonal sheets-in-cage (G) STEM image of the Ni-Mn-S

hybrid and the mapping results of Ni (G-a), Mn (G-b), and S (G-c) components.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

The dosage of Na2S triggered the formation of numerous tiny particles anchored onto the

surfaces of the sheets as indicated by the contrast difference (Fig. 3D&E) because of additional

nucleation. This implies that the surfaces of hexagonal sheets were used as firm scaffolds for

further nucleation during the anion exchange processes. These nucleation centers accelerated the

growth at planar and vertical directions, forming a sandwich-like architecture (Fig.3 D-E), which

Accepted Manuscript
is very favourable for charge storage because of establishing a large electroactive area, an

excellent porosity, and much accessible electrolyte at the top and core active sites. The HR-TEM

image of the Ni-Mn-S hybrid depicted in Fig. 3F shows well-resolved lattice fringes with

interplanar spacings of 0.334 nm and 0.392 nm, corresponding to [101] and [102] planes of

hexagonal NiS and γ-MnS, respectively. In addition, the EDS mappings (Fig. 3Ga-c) clearly

reveal a sustained homogeneous distribution of Ni, Mn, and S elements in the hybrid. The

distribution area of each component is highly overlapped, confirming the possible incorporation

of sulfur species in the whole matrix. No obvious elemental segregation was detected in the

formed heterostructure. The high doping of sulfur element can improve the cycle life of the

electrode material in an alkaline electrolyte [24a]. Moreover, the analysis demonstrates that the

molecular ratio of the Ni/Mn is around 1.32. This finding verifies that the hybrid Ni-Mn-S

sheets-in-cage can be successfully prepared when substitutional sulfur species integrated during

low-temperature sulfurization reaction. The high relative crystallinity and uniformity of

sulfurized product is mainly accredited to the crystallized and consolidated structure of the

layered precursor. Also, the adherence properties between the adjacent sheets are well preserved

without any detectable scattering. In spite of different reaction kinetics at both surface and core

whereas the anion exchange process is mostly accomplished on the outer surfaces, a unique

crystalline structure was achieved and thereby we can conclude that most of Ni and Mn occupies

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

the adequate position. Along the line, a self-doping conceptual can be speculated to well address

the phenomenon related to conductivity and metallic states.

Additionally, the fabrics of Ni-S and Ni-Mn-S based materials were further characterized by

BET technique as displayed in Fig. 4A-B. As we expected, the SBET and pore volume of the Ni-

Mn-S hybrid (86.8 m2 g–1 and 0.087 cm3 g–1, respectively) are much larger than those of

Accepted Manuscript
monometallic Ni-S (~50 m2 g–1 and 0.068 cm3 g–1, respectively). The larger surface area can

endow the hybrid with a higher electrochemical active surface area and a sufficient number of

electroactive sites at heterointerfaces during redox reactions. Moreover, the binary Ni-Mn-S

possesses abundant mesoporosity with a relatively high intense peaks between 2 and 25 nm,

which favors the diffusion of ions inside the open mesopores at low resistance. The possible

crystal defects occurred due to the coexistence of Ni-S and MnS phases might have a positive

role influencing the relatively large surface area of the composite.

Fig. 4 (A) Nitrogen adsorption−desorption isotherms of the prepared materials and (B) the

corresponding pore size distribution

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Electrocapacitive performances

In order to assess the electrochemical performance of binary Ni-Mn-S hexagonal sheets-in-cage

as electrode material for supercapacitors compared to individual component Ni-S and MnS,

cyclic voltammetry (CV), galvanostatic charge/discharge (GCD), and electrochemical impedance

Accepted Manuscript
spectroscopy (EIS) measurements were conducted in a 6 M KOH solution with a three-electrode

assembly. The typical CV curves of the electrodes at a scan rate of 10 mVs‒1 over a potential

window from ‒0.05 to 0.45 V are shown in Fig. 5A. Apparently the integrated area of the hybrid

Ni-Mn-S is extremely larger, which ascribed to the enhanced electrical conductivity and richer

electroactive sites of the fascinating electrode material. This implies an improved capacitive

activity and proves the fundamental roles of Ni and Mn species in a pseudocapacitive activity.

Furthermore, the redox current peaks through the anodic and cathodic scans observed for each

electrode are attributed to reversible reactions of Ni2+/Ni3+ and Mn2+/Mn3+ in an alkaline media

with the OH− assistance, suggesting an optimal battery-type Faradaic mode with respect to redox

reactions. The anodic and cathodic peaks of the Ni-Mn-S composite are found at similar

potentials compared to the literature [42]. Accordingly, the redox reactions can be addressed as

shown (equations 2&3):

OH‒+NiS ↔ e‒+NiSOH (1)

OH‒+MnS ↔ e‒+MnSOH (2)

A series of CV profiles (Figs. 5B&S5A-B) were recorded for the Ni-Mn-S, Ni-S, and MnS based

electrodes with various scan rates from 2 to 100 mVs−1. As shown, the peaks currents increase

with the increase of scanning rate and the redox peaks are observed clearly even at a high scan

rate of 100 mVs−1. Despite the shift of both anodic and cathodic peaks towards the higher

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

positive potential and lower negative potential directions, respectively, the features of the CV

curves demonstrates that the electrodes are preferable for rapid charge/discharge operations.

Impressively, the shape of the CV curves of the composite (Fig. 5B) was well maintained, while

the peaks currents improve, implying that the electrode has excellent rate performance that can

result in fast redox reactions. To figure out the advantages of hierarchical Ni-Mn-S composite,

Accepted Manuscript
we also tested the GCD responses compared with those of monometallic Ni-S (Fig. S5C) and

MnS (Fig. S5D). Fig. 5C displays the GCD curves of binary Ni-Mn-S and single component Ni-

S and MnS modified electrodes at a current density of 1 A g−1 over a potential range from ‒0.05

to 0.45 V. Apparently, the composite Ni-Mn-S electrode exhibits the largest discharge time and

achieves a high specific capacitance of 1664 F g −1 at 1 A g−1 compared to Ni-S (842 F g −1) and

MnS (584 F g −1). The specific capacitances of all electrodes at various current densities from 0.5

to 50 A g−1 are presented in Figs. 5D&E. The plateaus in the GCD curves are mainly originated

from the redox reactions. Moreover, the GCD curves of the electrodes are almost symmetrical,

demonstrating a good reversibility of the process and high-rate charge/discharge capacity [43],

which is accord with the CV spectra. As can be seen, the specific capacitances of binary Ni-Mn-

S and, bare Ni-S, and MnS based electrodes at a higher current density of 50 A g−1 are 785 F g−1,

307 F g−1, and 178.4 F g−1, respectively, corresponding to a capacitance retention of about

47.2%, 36.5%, and 30.5% respectively, compared to their specific capacitances at 1 A g−1. This

proves that the hybrid Ni-Mn-S delivers not only a large specific capacitance but also an

excellent capacitance retention at a high rate because of the firm combination of active materials.

The improvement in specific capacity of the hybrid electrode may be caused by the introduction

of the Mn ions, which promoted the number of valid ions from the Ni-S and bring the

complementary advantages between Ni and Mn species in the resulting composite. In fact, the

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

decrease in specific capacitance with increasing the current density might largely come from the

diffusion limitation and migration of electrolyte ions, which elevates the insufficiency of inner

active area for charge conservation at higher current densities.

The specific capacitances of the as-prepared electrodes are plotted as a function of current

density as displayed in Fig. 5E. Evidently, the capacitive behavior of the Ni-Mn-S electrode

Accepted Manuscript
outperforms that of pristine Ni-S and MnS based electrodes at all current densities. We speculate

that the enlarged surface area and pore volume of the composite can play a role contributing to

the improved capacitive activity. Besides, the synergy between the Ni and Mn species is more

beneficial to boost the capacitive performance and the specific capacitance of the composite is

originated from the pseudocapacitances of both electrochemically active components.

EIS measurements of the electrodes were carried out to reveal the intrinsic electrical conductivity

and convenient ion transfer/diffusion of the hybrid Ni-Mn-S. As presented in Fig. 5F, the

intersection with the x-axis at the high frequency region reflects the resistance to electrolyte (RS),

while the diameter of the semicircle denotes the charge transfer resistance (RCT). Additionally,

the straight line in the low frequency region illustrates the diffusion resistance of electrolyte ions

and carriers with the hierarchical electrode material. The adopted equivalent circuit model for the

fitted Nyquist plots is shown in Fig. S6. Obviously, the Ni-Mn-S composite displays a smaller

RCT value (2.88 Ω) than single component Ni-S (4.5 Ω) and MnS (6.29 Ω), indicating a more

flexible process of charge transportation at electrode/electrolyte heterointerfaces due to the low

intrinsic resistance of the active pseudocapacitive composite and small contact resistance of the

developed electrode. Interestingly, considering the reduction of RCT from Ni-Mn-S to Ni-S, it is

proved that Mn species could facilitate the charge transfer process. In addition, the solution

resistances of the Ni-Mn-S, Ni-S, and MnS were 0.37, 0.52, and 0.65 Ω, respectively. The

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

relatively larger slope of the low frequency line for the Ni-Mn-S electrode, manifesting a lower

ions diffusion resistance and thereby a higher energy storage behavior.

Accepted Manuscript
Fig. 5 (A) Comparison of the CV curves of the fabricated electrodes in a 6 M KOH solution at

10 mV s-1 scan rate. (B) CV curves of the mixed Ni-Mn-S electrode in a 6 M KOH at different

scanning rates. (C) Comparative GCD curves of the electrode at a current density of 1 A g−1. (D)

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

GCD curves of the hybrid Ni-Mn-S electrode at various current densities. (E) Plots of specific

capacitance as a function of current density. (F) Nyquist plots for the developed electrodes.

The long lifespan is of a great importance for constructing energy storage devices with desirable

efficiencies. The cycling stability of the various electrodes was evaluated by GCD test as

depicted in Fig. 6. It can be seen that the specific capacity of Ni-Mn-S composite is much higher

Accepted Manuscript
than that of monometallic Ni-S and MnS all over the 3000 cycles of reiterated charge/discharge

processes. The specific capacity retentions are 90.7%, 81%, and 68.9% for the Ni-Mn-S hybrid,

Ni-S, and MnS, respectively. Thus, the Ni and Mn ions are beneficial for superior cycling

performance. The slight shift of the GCD curves of the first 10 cycles and last 10 cycles for the

hybrid Ni-Mn-S compared to monometallic Ni-S and MnS suggests a good cycling stability of

the hybrid supercapacitor. SEM images of the Ni-Mn-S electrode after 3000 cycles at 10 A g−1 are

illustrated in Fig. S7. It is apparent that the surface morphologies remain unchanged after cycling at

such a high rate, confirming the extraordinary rate performance and long cycle lifetime of the

electrode. It can be deduced that the hexagonal sheets-in-cage-like morphologies serve as strong

supporting frameworks for small and interconnected nanoparticles to create a hierarchical

micro/nanostructure with enhanced mechanical and chemical stabilities.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Accepted Manuscript
Fig. 6 Capacity retention for 3000 cycles at a current density of 10 A g−1 for Ni-Mn-S (A), Ni-S

(B), and MnS based electrode (C). The insets illustrate the GCD plots for the first 10 cycles and

last 10 cycles of the electrodes during the charge/discharge processes.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

To further explore the energy storage capacity of heterostructured Ni-Mn-S composite for

practical use, an asymmetric supercapacitor device was established in a 6 M KOH solution

according to the literature [44]. The CV curves of activated carbon (AC) and Ni-Mn-S modified

electrodes with a three-electrode implementation at a scan rate of 10 mVs−1 are displayed in Fig.

7A. Referring to the reported data, the AC can achieve the double-layer capacitance features in a

Accepted Manuscript
KOH solution with the potential window from 0 to ‒1 V [23,45]. The corresponding GCD

profile of AC/NF shown in Fig. S8 reveals a triangle-like shape, implying the occurrence of

electrochemical double-layer capacitance properties and provides a specific capacitance of 175 F

g−1 at 1 A g−1. Thus, the total operating potential of the fabricated device is anticipated to reach

1.5 V. To attain an optimal electrochemical behaviour, the ratio of the positive to negative

electrode was adopted based on this relation:

ΔQPositive=ΔQNegative (3)

The mass ratio of the active Ni-Mn-S and the AC was adjusted to approximately be 0.21 based

on their specific capacitance and corresponding potential windows. To extend the operational

voltage window of the as-assembled Ni-Mn-S//AC asymmetric device, a total potential of 1.5 is

desirable to fully utilize the different potential ranges of both integrated electrodes. Fig. 7B

shows the typical CV curves of the Ni-Mn-S//AC asymmetric device with various scanning rates

from 5 to 100 mV s-1. Such CV responses might probably due to the dual effect of electric

double-layer capacitance and pseudocapacitance of the positive electrode within the test window.

The well-preserved shape proves the exceptional reversibility of the constructed device. Next,

the GCD curves were evaluated at various current densities as shown in Fig. 7C. The limited

plateaus in the analyzed spectra suggest that water splitting processes are not visible. On a hand,

the shape and symmetry of the GCD curves maintained similar, confirming an excellent

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

reversibility. Fig. 7D displays the coulombic efficiency of the device at different current

densities. The good coulombic efficiency clearly suggests the splendid rates of charge/discharge

reactions. The corresponding specific capacitances analyzed from the GCD curves are plotted in

Fig. 7E. The constructed device delivers a specific capacitance of 129.4 F g-1 at 1 A g-1 and

conserves around 48.5% (~ 62.8 F g-1) at 20 A g-1, suggesting the superior rate capability of the

Accepted Manuscript
device. Energy and power densities are significant indicators to well-assess the capability of the

device for charge storage. A Ragone plot (Fig. 7F) was performed for the Ni-Mn-S//AC

asymmetric capacitor to address the achieved energy and power densities as expressed by the

following relations (5&6):

(∆𝑉)2
𝐸= ×𝐶 (4)
2

𝐸
𝑃 = ∆𝑡 (5)

Where E and P are the energy density (Wh kg−1) and power density (W kg−1), respectively,

while, C and Δt represent the cell capacity (F g-1) and discharge time (s), respectively. The power

density increases with the decrease of the energy density, since large charges can be conserved

into the cell at a slight accumulation rate. The Ni-Mn-S//AC asymmetric device shows a

maximum energy density of 40.4 Wh kg−1 and a power density of 750 W kg−1 at 1 A g−1. Even at

a high discharge current of 20 A g−1, the energy density could reach ~ 19.6 Wh kg−1 at a high-

power density of 15 kW kg−1. The energy and power densities for the proposed devise are highly

competitive to those obtained by some previously reported asymmetric capacitors devices

established with earth abundant elements and carbon materials as the positive and negative

electrode, respectively, such as α-MnS/N-rGO//N-rGO (27.7 W h kg‒1 at 800 W kg‒1 and 16.1 W

h kg‒1 at 20 kW kg‒1) [51a], Ni‒Mn sulfides//RGO (36 W h kg‒1 at 775 W kg‒1 and 16.9 W h

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

kg‒1 at 15.5 kW kg‒1) [51b], Ni0.67Co0.33MoO4//RGO (25.6 Wh kg‒1 at 775 Wkg‒1 and 13.2 Wh

kg‒1 at 7.75 kW kg‒1) [51c], NiCo2S4@NiCo2S4//rGO (24.9 Wh kg‒1 at 334 W kg‒1 and 12.6 Wh

kg‒1 at 2.44 kW kg‒1) [51d], Co2.5Mn0.5 sulfide//RGO (22.3 Wh kg‒1 at 750 W kg‒1 and 11 Wh

kg‒1 at 30 kW kg‒1) [52a], NiS2//rGO (32.76 W h kg‒1 at 954 W kg‒1 and 11.19 W h kg‒1 at 1.35

kW kg‒1) [52b], CMTs-1000/Ni2CoS4//AC (28.1 Wh kg−1 at 753 Wkg−1 and 17.7 Wh kg−1 at

Accepted Manuscript
17.2 kW kg−1) [52c], MnO2//Bi2O3 (11.3 Wh kg−1 at 352.6 Wkg−1 and 9.1 Wh kg−1 at 3.37 kW

kg−1) [53a], K0.27MnO2·0.6H2O//AC (25.3 Wh kg−1 at 140 Wkg−1 and 17.6 Wh kg−1 at 2.0 kW

kg−1) [53b], PB@MnO2//PG (16.5 Wh kg−1 at 550 Wkg−1 and 11.3 Wh kg−1 at 5.497 kW kg−1)

[53c], Ni3S2@MoS2//rGO (21.7 Wh kg−1 at 400 Wkg−1 and 12 Wh kg−1 at 2.40 kW kg−1) [54a],

and NiCo2S4//AC (27.5 Wh kg−1 at 747 Wkg−1 and 18.75 Wh kg−1 at 6.03 kW kg−1) [54b].

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Accepted Manuscript
Fig. 7 (A) Comparative CV curves for the hybrid Ni-Mn-S and AC based-electrode at 10 mV s-1

with a three-electrode measurements, (B) CV curves of the as-designed Ni-Mn-S//AC

asymmetric device recorded with various sweep rates, (C) GCD curves for the device at different

current densities, (D) coulombic efficiency of the assembled device at different current densities,

(E) relationship of specific capacitance of the device versus the current density, and (F) Ragone

plot of the developed Ni-Mn-S//AC device compared with some existed results.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Another critical aspect to evaluate the feasibility an asymmetric device for the energy storage is

the long-term sustainability. Herein, the cycling stability of the optimized device was studied

using GCD technique at a high current density of 10 A g−1 for 10,000 successive cycles as shown

in Fig. 8. The capacity retention reaches ⁓85.5% with a near offset of the corresponding GCD

curves for the first and last cycles, benefiting from the engineered hierarchical architecture and

Accepted Manuscript
unique chemical composition. The superior stability can be credited to the smaller RCT which

allows elastic redox reactions at interfaces. It is well known that Ni-S would be transformed into

Ni(OH)2 during cycling in a basic environment, resulting in a modest retention of specific

capacity [55,56a]. This implies that the Mn cations act like strong matrices and adsorbs the

generated stresses during electrochemical reactions, leading to an excellent cycling retention.

Such cycling performance is extraordinary for pseudocapacitance-based asymmetric

supercapacitors.

Fig. 8 Cyclic stability of the fabricated Ni-Mn-S//AC capacitor device at 10 A g−1

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

To sum up, the intrinsic charge storage performance of the synthesized Ni-Mn-S hexagonal

sheets-in-cage (Tables S2&S3) is a result of the composition, surface engineering, and

microstructure as well.

a) The 3D interconnected sheets-in-cage structures with a relatively large surface area and a

smart porous texture offer sufficient connections at the active material/electrolyte

Accepted Manuscript
heterointerfaces and stimulate the overall utilization of the electrode.

b) The formation of richer void spaces into the hierarchal heterostructures allows larger

contact areas and can directly bring the internal active surface layers into contact with

electrolyte ions and hence minimize the ion diffusion pathways [56b,56c].

c) The hexagonal sheets-like matrices are closer enough with a preferable mechanical

strength and interfacial vacancies not only facilitate the mass and electron diffusion

during electrochemical reaction process but alleviate the stress developed by sudden

volume changes.

d) The strong and solid interaction between building blocks enables an extraordinary

structure stability at higher current densities and thus resulting in a high rate capability

without damaging their substantial and flexible architecture.

e) The synergetic effect of Ni-S and MnS components with multiple oxidation states also

affords a positive influence on the electrochemical properties and improves the total

capacitance throughout providing abundant electroactive sites and faster electronic

conductivity for Faradaic redox reactions than monometallic sulfides due to a smaller

band gap and a rapid electron transfer [13,23,57].

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Conclusion

Based on this work, 3D hierarchically porous Ni-Mn-S hexagonal sheets-in-cage heterostructures

were designed by a facile hydrothermal approach assisted anion-exchange reactions as an

effective electrode material for supercapacitor applications. Of particular interest is the fact that

Accepted Manuscript
the rationally fabricated hybrid substantiated the advantages of collaborative effect of

exceptional pseudocapacitive components and fast diffusion/migration of electrolyte ions

through the whole structure during charge/discharge reactions. The Ni-Mn-S complex displayed

a high specific capacitance of 1664 Fg‒1 at 1 A g‒1 and it’s corresponding asymmetric

supercapacitor device with a potential window of 1.5 V exhibited a maximum energy density of

40.4 Wh kg−1 at a power density of 750 W kg−1. The assembled Ni-Mn-S//AC asymmetric device

showed an excellent cycling stability with a retention of 85.5% after 10,000 cycles of repeated

charge/discharge processes. The rational design and abundant availability of ingredients suggest

that scalable quantities of the hybrid Ni-Mn-S may be economically fabricated and applied as

efficient electrode material for renewable energy technologies in the near future.

Acknowledgements
This work is supported by national key research and development program (Grant No.

2016YFB0901600), Zhejiang Provincial Natural Science Foundation of China under Grant No.

LY19E020014, and NSFC (Grant No. 21303162 and Grant No. 11604295).

Experimental

Synthesis

The composite Ni-Mn-S was fabricated by a simple two‐step hydrothermal route assisted anion

exchange reaction. In a typical synthesis strategy, 1 mM nickel sulfate hexahydrate

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

(NiSO4.6H2O), 1 mM manganese sulfate monohydrate (MnSO4.xH2O), 3 mM ammonium

fluoride (NH4F), and 1 mM urea (CH4N2O) were mixed in 27 mL double distilled water under a

vigorous magnetic stirring for 1 h. Then, the resultant mixture was loaded into a Teflon‐lined

stainless-steel autoclave with a total capacity of 40 mL, sealed, and kept in an oven at 180 °C for

12 h before cooling down to room temperature. The sample was collected by centrifugation,

Accepted Manuscript
washed thoroughly with deionized and absolute alcohol, and dried overnight at 60 °C. After that,

the as-obtained precursor was dispersed into 30 mL deionized water under stirring and thereafter

2 mM of Na2S·9H2O was added. After being stirred by 30 min, the mixture was sealed into a

Teflon-lined autoclave and treated thermally at 180 oC for 12 h. The formed black yield was

separated and rinsed several times with distilled water and ethanol before being dried at 60 oC in

a vacuum oven for 8 h. Moreover, pristine Ni-S and MnS were also prepared by same synthesis

route.

Materials characterization

The microstructural properties and surface morphologies were characterized by a scanning

electron microscopy (SEM, Hitachi S4800) and a field-emission transmission electron

microscopy (FE-TEM, FEI Tecnai G2 F20) equipped with energy-dispersive X-ray spectrometry

(EDS). The crystal structure of the prepared materials was examined by an X-ray diffraction

(XRD, Shimadzu-6000) with Cu Kα radiation (λ=0.15 nm) in the 2θ range from 10° to 80°. The

oxidation states of elements were analyzed by an X-ray photoelectron spectroscopy (XPS,

Thermo Fisher Scientific ESCALAB 250Xi) with an Al Kα radiation source. The specific

surface area and pore size properties were evaluated by liquid gas adsorption-desorption

measurements using Autosorb (Quantachrome Instruments) at 77 K. Before collecting the

isotherms, the samples were treated thermally for 6 h at 200 °C with a N2 atmosphere. The

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

specific surface area (SBET) was assessed according to the Brunauer–Emmett–Teller (BET)

method, while the pore size distribution investigated by using Barrett–Joyner–Halenda (BJH).

Electrochemical measurements

The electrochemical properties of the as-prepared materials were conducted on a CHI660E

Accepted Manuscript
electrochemical instrument in a 6 M KOH solution at room temperature. A platinum plate and

Hg/HgO were employed as the counter and reference electrode, respectively, while an acidic

treated Ni-substrate covered with a uniform film of the active material was employed as the

working electrode. The slurry was prepared by mixing the active material, carbon black, and

polyvinylidene fluoride (PVDF) with a ratio of 8:1:1 (w/w) into N-methyl pyrrolidone (NMP)

solvent. The resulting slurry was carefully deposited onto the current collector (1 cm*1 cm),

pressed with a pressure of ⁓20 MPa, and then dried in a vacuum oven at 60 °C to for 8 h to

achieve an active loading of ⁓1.5 mg cm‒2. The specific capacitance (C, F g-1) of the electrode

was measured according to the following equation.

I∗t
C = ΔV∗m (6)

Where I and t are the constant discharge current (A) and discharge time (s), respectively, while

ΔV and m represent the potential window (V) and active material loading (g), respectively. An

asymmetric supercapacitor was fabricated based on the hybrid Ni-Mn-S as the positive electrode

and conventional AC as the negative electrode with a thin porous membrane of cellulose paper.

The CV and GCD characterizations were performed in the potential window from -0.05 V to

0.45 V for the three-electrode measurements and from 0 V to 1.5 V in case of asymmetric

supercapacitor experiment. Moreover, the EIS profiles were collected with a frequency range

from 100 kHz to 0.01 Hz.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

References
1. a) T. Lin, I.-W. Chen, F. Liu, C. Yang, H. Bi, F. Xu, F. Huang, Science 2015, 350, 1508-
1513; b) D. Khalafallah, M. Zhi, Z. Hong, Top Curr Chem. 2019, 377, 29; c) D.
Khalafallah, C. Ouyang, M. Zhi, Z. Hong, ChemElectroChem. 2019, 6, 5191–5202; e) D.
Khalafallah, N. Akhtar, O.Y. Alothman, H. Fouad, K.A. Khalil, Solid State Sci., 2017,
71, 51-60.
2. a) A. S. Arico, P. Bruce, B. Scrosati, J. M. Tarascon, W. Van Schalkwijk, Nat. Mater.
2005, 4, 366–377; b) D. Khalafallah, O. Y. Alothman, H. Fouad, K. A. Khalil, Int. J.
Hydrogn. Energy 2018, 43, 2742-2753.

Accepted Manuscript
3. S. Peng, L. Li, H.B. Wu, S. Madhavi, X.W. Lou, Adv. Energy Mater. 2015, 5, 1401172.
4. Y. P. Huang, F. Cui, M. Q. Hua, L. Xu, Y. Zhao, J. B. Lian, J. Bao, H. M. Li, Chem.
Asian. J 2018, 13, 3212-3221.
5. W. Tian, X. Wang, C. Zhi, T. Zhai, D. Liu, C. Zhang, D. Golberg, Y. Bando, Nano
Energy 2013, 2, 754-763.
6. S. H. Liu, G. H. Xu, J. Li, B. Wang, Z. Y. Huang, Q. Chen, X. Qi, ChemElectrochem
2018, 5, 2250-2255.
7. K. Krishnamoorthy, P. Pazhamalai, G.K. Veerasubramani, S.J. Kim, J. Power Sources
2016, 321, 112-119.
8. K. Krishnamoorthy, S. Thangavel, J. Chelora Veetil, N. Raju, G. Venugopal, S.J. Kim,
Int. J. Hydrogen Energy 2016, 41, 1672-1678.
9. C. Liu, F. Li, L.-P. Ma, H.-M. Cheng, Adv. Mater. 2010, 22, E28–E62.
10. G. Wang, L. Zhang, J. Zhang. Chem. Soc. Rev. 2012, 41, 797–828.
11. M. Acerce, D. Voiry, M. Chhowalla, Nat. Nanotechnol. 2015, 10, 313-318.
12. J. Tang, Y. Ge, J. Shen, M. Ye. Chem. Commun. 2016, 52, 1509–1512.
13. S. Liu, S. C. Jun. J. Power Sources 2017, 342, 629-637.
14. X. Li,Q. Li, Y. Wu, M. Rui, H. Zeng. ACS Appl. Mater. Interfaces 2015, 7,
19316−19323.
15. S. Sahoo, R. Mondal, D. J. Late, C. S. Rout. Microporous Mesoporous Mater. 2017, 244,
101-108.
16. J. Xiao, L. Wan, S. Yang, F. Xiao, S. Wang. Nano Lett. 2014, 14, 831–838.
17. H. Zhang, X. Yu, D. Guo, B. Qu, M. Zhang, Q. Li, T. Wang, ACS Appl. Mater. Interfaces
2013, 5, 7335.
18. G. Yu, L. Hu, N. Liu, H. Wang, M. Vosgueritchian, Y. Yang, Y. Cui, Z. Bao, Nano Lett.
2011, 11, 4438
19. J. Cheng, H. Yan, Y. Lu, K. Qiu, X. Hou, J. Xu, L. Han, X. Liu, J.-K. Kim, Y. Luo, J.
Mater. Chem. A 2015, 3, 9769-9776.
20. L. Shen, L. Yu, H.B. Wu, X. Yu, X. Zhang, X.W. Lou, Nat. Commun. 2015, 6, 6694.
21. H. Chen, J. Jiang, L. Zhang, H. Wan, T. Qi, D. Xia. Nanoscale 2013, 5, 8879–8883.
22. J. Xiao, L. Wan, S Yang, F. Xiao, S. Wang. Nano Lett. 2014, 14, 831–838.
23. H. Peng, G. Wei, K. Sun, G. Ma, E. Feng, X. Yang, Z. Lei, New J. Chem., 2018, 42,
18328—18334.
24. a) J. Balamurugan, C. Li, T. D. Thanh, O.-K. Park, N. H. Kim, J. H. Lee, J. Mater. Chem.
A, 2017, 5, 19760-19772; b) H. Wan, J. Liu, Y. Ruan, L. Lv, L. Peng, X. Ji, L. Miao, J.
Jiang, ACS Appl. Mater. Interfaces 2015, 7, 15840−15847; c) Z. Chen, Z. Wan, T. Yang,
M. Zhao, X. Lv, H. Wang, X. Ren, X. Mei, Sci. Rep., 2016, 6, 25151; d) J. Li. W. Xu. J.
Luo. D. Zhou. D. Zhang. L. Wei. P. Xu, D. Yuan, Nano-Micro Lett., 2018, 10, 6; e) W.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

Li, B. Zhang, R. Lin, S. Ho-Kimura, G. He, X. Zhou, J. Hu, I. P. Parkin, Adv. Funct.
Mater. 2018, 28, 1705937.
25. a) L. Mi, W. Wei, Z. Zheng, G. Zhu, H. Hou, W. Chen, X. Guan, Nanoscale 2014, 6,
1124; b) Z.-D. Huang, K. Zhang, T.-T. Zhang, X. Li, R.-Q. Liu, X.-M. Feng, Y. Li, X.-J.
Lin, Y.-B. He, X.-S. Yang, Y.-W. Ma, J. Mater. Chem. A 2015, 3, 20886-20891.
26. Z.-D. Huang, K. Zhang, T.-T. Zhang, R.-Q. Liu, X.-J. Lin, Y. Li, T. Masese, X. Liu, X.-
M. Feng, Y.-W. Ma, Energy Storage Mater. 2016, 5, 205-213.
27. Y. Chen, B. Qu, L. Hu, Z. Xu, Q. Li, T. Wang, Nanoscale 2013, 5, 9812-9820.
28. G. Cheng, Y. Yan, R. Chen. New J. Chem., 39, 2015, 676-682.

Accepted Manuscript
29. R. Zou, Z. Zhang, M.F. Yuen, M. Sun, J. Hu, C.-S. Lee, W. Zhang, NPG Asia Mater.
2015, 7, e195.
30. a) H. Huo, Y. Zhao, C. Xu, J. Mater. Chem. A 2014, 2, 15111-15117; b) K.
Krishnamoorthy, G. K. Veerasubramani, S. Radhakrishnan, S. J. Kim, Chem. Eng. J.
2014, 251, 116-122.
31. a) B. Guan, Y. Li, B. Yin, K. Liu, D. Wang, H. Zhang, C. Cheng, Chem. Eng. J., 2017,
308, 1165–1173; b) X. Yang, L. Zhao, J. Lian, J. Power Sources, 2017, 343, 373–382.
32. a) X. Li, J. Shen, N. Li, M. Ye, J. Power Sources 2015, 282, 194–201; b) G. Li, B. He,
M. Zhou, G. Wang, N. Zhou, W. Xu, and Z. Hou. ChemElectroChem 2017, 4, 81 – 89.
33. J. Y. Liao, D. Higgins, G. Lui, V. Chabot, X. C. Xiao, Z. W. Chen, Nano Lett. 2013, 13,
5467.
34. S. D. Perera, M. Rudolph, R. G. Mariano, N. Nijem, J. P. Ferraris, Y. J Chabal, K. J.
Balkus Jr., Nano Energy 2013, 2, 966.
35. J. H. Zhong, A. L. Wang, G. R. Li, J. W. Wang, Y. N. Ou, Y. X. Tong, J. Mater. Chem.
2012, 22, 5656; b) D. Khalafallah, L. Xiaoyu, M. Zhi, Z. Hong, ChemElectroChem
doi.org/10.1002/celc.201901423; c). D. Khalafallah, M. Zhi, Z. Hong, Energy Technol.
2019, 7, 1900548.
36. X. Zhang, Z. Xing, L. L. Wang, Y. C. Zhu, Q. W. Li, J. W. Liang, Y. Yu, T. Huang, K.
B. Tang, Y. T. Qian and X. Y. Shen, J. Mater. Chem., 2012, 22, 17864–17869.
37. A. Pramanik, S. Maiti, S. Mahanty, J. Mater. Chem. A 2014, 2, 18515.
38. Y. Zhang, Y. V. Lim, S. Huang, M. E. Pam, Y. Wang, L. K. Ang, Y. Shi, H. Y. Yang.
Small 2018, 14, 1800898.
39. Y. Xiang, Z. Yin, X. Li, J Solid State Electrochem., 2014, 18, 2123.
40. a) L. Pan, J. J. Zou, S. B. Wang, X. Y. Liu, X. W. Zhang, L. Wang, ACS Appl. Mater.
Interfaces 2012, 4, 1650; b) W. Wei, L. Mi, Y. Gao, Z. Zheng, W. Chen, X. Guan, Chem.
Mater. 2014, 26, 3418−3426.
41. L. Shen, L. Yu, H. Wu, X. Yu, X. Zhang, X. W. Lou, Nat. Commun., 2015, 6, 6694.
42. J. Zhao, J. Chen, S. Xu, M. Shao, Q. Zhang, F. Wei, J. Ma, M. Wei, D. G. Evans, X.
Duan, Adv. Funct. Mater. 2014, 24, 2938.
43. H. Chen, J. Jiang, L. Zhang, H. Wan, T. Qi, D. Xia, Nanoscale 2013, 5, 8879.
44. W. Wang, N. Zhang, Z. Shi, Z. Ye, Q. Gao, M. Zhi and Z. Hong, Chem. Eng. J. 2018,
338, 55.
45. P.-Y. Lee, L.-Y. Lin, J Colloid Interface Sci., 2019, 538, 297–307.
46. Z. Jiang, W. Lu, Z. Li, K. H. Ho, X. Li, X. Jiao, D. Chen, J. Mater. Chem. A 2014, 2,
8603.
47. Y. Li, L. Cao, L. Qiao, M. Zhou, Y. Yang, P. Xiao, Y. Zhang, J. Mater. Chem. A, 2014,
2, 6540.

This article is protected by copyright. All rights reserved.


Chemistry - A European Journal 10.1002/chem.201904991

48. Y. Zhu, Z. Wu, M. Jing, X. Yang, W. Song and X. Ji, J. Power Sources 2015, 273,
584−590.
49. H. Chen, J. Jiang, L. Zhang, D. Xia, Y. Zhao, D. Guo, T. Qi, H. Wan, J. Power Sources
2014, 254, 249.
50. W. Hu, R. Chen, W. Xie, L. Zou, N. Qin, D. Bao, ACS Appl. Mater. Interfaces, 2014, 6,
19318.
51. a) H. Y. Quan, B. C. Cheng, D. Z. Chen, X. H. Su, Y. H. Xiao, S. J. Lei, Electrochim
Acta 2016, 210, 557–566; b) J. Cao, S. Yuan, H. Yin, Y. Zhu, C. Li, M. Fan, H. Chen. J
Solgel Sci Technol., 2018, 85, 629–637; c) H. C. Chen, S. Chen, Y. Y. Zhu, C. Li, M. Q.

Accepted Manuscript
Fan, D. Chen, G. L. Tian, K. Y. Shu, Electrochim Acta 2016, 190, 57–63; d) H. C. Chen,
S. Chen, H. Y. Shao, C. Li, M. Q. Fan, D. Chen, G. L. Tian, K. Y. Shu, Chem. Asian. J.,
2016, 11, 248–255.
52. a) S. Chen, H. C. Chen, C. Li, M. Q. Fan, C. J. Lv, G. L. Tian, K. Y. Shu, J. Mater Sci.,
2017, 52, 6687–6696; b) Y. J. Ruan, J. J. Jiang, H. Z. Wan, X. Ji, L. Miao, L. Peng, B.
Zhang, L. Lv, J. Liu, J Power Sources 2016, 301,122–130; c) K. Wang, R. Yan, X. Tian,
Y. Wang, S. Lei, X. Li, T. Yang, X. Wang, Y. Song, Y. Liu, Z. Liu, Q. Guo.
Electrochimica Acta 2019, 302, 78-91.
53. a) H. H. Xu, X. L. Hu, H. L. Yan, Y. M. Sun, Y. H. Huang, Adv. Energy Mater., 2015, 5,
201401882; b) Q.T. Qu, L. Li, S. Tian, W.L. Guo, Y.P. Wu, R. Holze, J. Power Source
2010, 195, 2789-2794; c) A. K. Das, R. Bera, A. Maitra, S. K. Karan, S. Paria, L. Halder,
S. K. Si, A. Bera, B. B. Khauta, J. Mater. Chem. A 2017, 5, 22242-22254.
54. a) L. Huang, H. Hou, B. Liu, K. Zeinu, X. Zhu, X. Yuan, X. He, L. Wu, J. Hu, J. Yang,
Appl. Surf. Sci. 2017, 425, 879-888; b) L. Liu, T. Chen, H. Rong, Z. Wang, J. Alloys.
Compd., 2018, 766, 149-156.
55. J. Yang, X. Duan, Q. Qin, W. Zheng, J. Mater. Chem. A, 2013, 1, 7880.
56. a) B. T. Zhu, Z. Wang, S. Ding, J. S. Chen, X. W. Lou, RSC Adv., 2011, 1, 397; b) L. G.
Beka, X. Li, W. Liu, Sci. Rep. 2017, 7, 2105; c) D. Khalafallah, O.Y. Alothman, H.
Fouad, K.A. Khalil, J. Electroanal. Chem. 2018, 809, 96-104.
57. a) Y. Zhao, Z. Shi, H. Li, C.-A. Wang, J. Mater. Chem. A, 2018, 6, 12782-12793; b) X.
Wang, Q. Zhang, J. Sun, Z. Zhou, Q. Li, B. He, J. Zhao, W. Lu, C.-p. Wong, Y. Yao, J.
Mater. Chem. A, 2018, 6, 8030–8038.

This article is protected by copyright. All rights reserved.

You might also like