You are on page 1of 26

Journal Pre-proofs

Revealing the role of crystal structure to catalysis: Inverse spinel phase Co-
Mn-based catalyst for Li-S batteries

Xuelin Huang, Peng Zeng, Yunfeng Lu, Juan Yang, Manfang Chen, Hong
Liu, Xianyou Wang

PII: S1385-8947(24)01977-6
DOI: https://doi.org/10.1016/j.cej.2024.150490
Reference: CEJ 150490

To appear in: Chemical Engineering Journal

Received Date: 12 December 2023


Revised Date: 14 February 2024
Accepted Date: 15 March 2024

Please cite this article as: X. Huang, P. Zeng, Y. Lu, J. Yang, M. Chen, H. Liu, X. Wang, Revealing the role of
crystal structure to catalysis: Inverse spinel phase Co-Mn-based catalyst for Li-S batteries, Chemical Engineering
Journal (2024), doi: https://doi.org/10.1016/j.cej.2024.150490

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2024 Published by Elsevier B.V.


Revealing the role of crystal structure to catalysis: inverse spinel
phase Co-Mn-based catalyst for Li-S batteries
Xuelin Huanga, Peng Zenga,b, Yunfeng Lua, Juan Yanga, Manfang Chena, Hong Liua, Xianyou Wanga

(a: National Base for International Science & Technology Cooperation, National Local Joint Engineering
Laboratory for Key Materials of New Energy Storage Battery, Hunan Province Key Laboratory of Electrochemical
Energy Storage & Conversion, School of Chemistry, Xiangtan University, Xiangtan 411105, China

b: Key Laboratory for Theoretical Organic Chemistry and Functional Molecules, Ministry of Education, Hunan
University of Science and Technology, Xiangtan 411201, China)

Abstract:

Li-S batteries are a promising energy storage system because of their inexpensive
price and high theoretical energy density (2600 Wh kg-1). However, the actual
performance of Li-S batteries is still hampered by the severe shuttle effect and the slow
reaction kinetics of lithium polysulfides (LiPSs). Balancing the high energy density and
long cycle life of Li-S batteries still face many challenges. In this work, inverse spinel
phase Co-Mn bi-metal oxides are proposed as an advanced sulfur reduction reaction
(SRR) catalyst to prevent irreversible loss of sulfur species and accelerate the reaction
kinetics of Li-S batteries. It has been found that both normal spinel phase Co2MnO4 (n-
CMO) and inverse spinel phase CoMn2O4 (i-CMO) have significant catalytic activity
to the conversion reaction of sulfur species. In the same time, based on the
electrochemical impedance spectroscopy (EIS) in variable temperature and in situ
ultraviolet visible (Uv-vis) spectroscopy, it has also been found that the i-CMO catalyst
shows much better performance than n-CMO since it could reduce the activation energy
of SRR reaction and promote the dissociation of the S8 ring. As a result, the i-CMO/S
cathode delivers a high initial discharge specific capacity oof 1386 mAh g-1 at 0.1 C
together with a low-capacity fading rate of 0.11% per cycle within 400 cycles. Besides,
when the i-CMO nanoparticles are loaded on the surface of carbon cloth (CC), the
CC@i-CMO/S cathode provides a high areal capacity of 4.26 mAh cm-2 at 0.1 C, in
which sulfur areal loading is 3.54 mg cm-2. Therefore, this study is a positive attempt
to study the relationship between catalytic performance and the crystal structure of the
materials, which will be conducive to the practical use of Li-S batteries.

Key words: inverse spinel phase; Co-Mn bi-metal oxides; Sulfur reduction reaction
catalyst; Reaction kinetics; Li-S batteries

Corresponding author: Xianyou Wang, Tel: +86 0731 58293377; E-mail: wxianyou@yahoo.com,
xywang@xtu.edu.cn
1
1. Introduction

As an electrochemical energy storage, lithium-ion batteries are playing an


important role in promoting human society to move forward in the direction of high
efficiency and convenience[1, 2]. Current commercial Li-ion batteries, however,
confront numerous obstacles in achieving the demands for high energy density, low
cost, and safety[3]. In this perspective, Li-S batteries stand out owing to their high
energy density (2600 Wh kg-1), low cost and high theoretical specific capacity (1675
mAh g-1). Furthermore, element sulfur is abundant in nature and non-hazardous to the
environment, making it a prospective high energy density storage system[4].
Unfortunately, despite the excellent features of Li-S batteries, their commercialization
still faces many obstacles, such as the poor conductivity of S and Li2S[5], the severe
shuttle effect of soluble lithium polysulfides (LiPSs)[6], and the slow reaction
kinetics[7]. In the discharge process of Li-S batteries, the long-chain LiPSs are formed
and tend to dissolve in the ether electrolyte, resulting in the passivation of the lithium
anode and the loss of the active material[8]. More seriously, the conversion of LiPSs
must overcome a high reactive energy barrier, so the LiPSs cannot be fully converted
in time, leading to an unsatisfied electrochemical performance.

Recently, numerous significant efforts have been made to enhance the


performance of Li-S batteries, including the development of cathode materials with
high S utilization and conductivity, separator modification, electrolyte improvement,
and lithium anode protection[8, 9]. For the cathode portion, mixing the active sulfur
with a variety of highly conductive cathode materials has been generally accepted by a
wide range of researchers[10, 11]. Among them, in the past decade or so, the conductive
carbon materials such as carbon nanotubes[12] and graphene[13] have attracted the
much more attention of researchers because of their high electrical conductivity[14].
However, the weak interaction force between the conductive carbon materials and
LiPSs is insufficient to prevent the shuttle effect of LiPSs intermediate[15, 16]. In this
regard, the research focus gradually shifted to the development of catalyst, which can
improve the reaction kinetics of the conversion reaction between LiPSs intermediates
and Li2S[17, 18]. Such an idea is considered to be a breakthrough in effectively solving
the shuttle effect[19]. In order to further improve the adsorption capacity of the LiPSs
intermediate and promote redox kinetics, researchers have begun to conduct a lot of
studies on transition metal compounds, such as metal oxides[20], metal carbides[21],
and metal sulfides[22]. These metal oxides can not only promote the binding of polar
LiPSs, but also significantly enhance the reaction kinetics of the conversion reaction,
thus promoting performances of Li-S batteries.

In lithium-sulfur batteries, sulfur reduction is the process during battery discharge


where lithium ions can be reduced polysulfide to more affordable sulfide[23, 24]. This

2
reaction is usually a key electrochemical process in lithium-sulfur batteries[25]. Metal
oxides play a catalytic role by forming metal-sulfur bonds, which weaken the S-S bonds
in LiPSs. This weakening facilitates the breaking and transformation of polysulfides,
promoting their conversion and preventing the shuttle effect[9, 26].

Polymetal oxides have shown promise in enhancing reaction kinetics and


electrochemical properties by providing additional active sites to capture
polysulfides[27, 28]. Moreover, transition metal oxides also possess distinctive
catalytic activity, particularly in supercapacitors and fuel cells, contributing to their
effectiveness in these applications[29]. Spinel-type transition metal oxides (STMOs)
are more widely employed in Li-S batteries than typical metal oxides due to their
distinct characteristics[24, 30]. On the one hand, the polar affinity of STMOs (AB2O4:
A and B represent divalent and trivalent metal ions, respectively)[31] allows it to adsorb
LiPSs more efficiently, thus inhibiting the shuttle effect[32]. On the other hand, it can
also increase the comprehensive performance by promoting the reaction kinetics of the
LiPSs conversion process[33]. Based on the above considerations, the spinel-type
NiCo2O4 catalyst[34] are developed and applied in Li-S batteries. Furthermore, Lin et
al. proposed that the highly polar NiCo2O4 can capture LiPSs via Ni-S and Co-S bonds,
thus increasing LiPSs conversion and alleviating the shuttle effect[35]. Similarly,
NiCo2O4[34] 、 NiFe2O4[36] and NiMoO4[37] are successively applied in research
work related to Li-S batteries. As far as we are concerned, in-depth reports on Co-Mn
bimetallic oxide catalysts are still scarce, especially the connection between crystal
configuration and catalytic activity.

In fact, Co-Mn bimetallic oxides are frequently applied in catalytic fields. For
example, Wu et al. found that when the normal spinel structure {Co}[FeCo]O4 is
transformed into the inverse spinel structure, the catalytic performance can be
significantly improved[38]. Subsequently, Gong et al. studied the application of inverse
spinel type Co[Co, Fe]O4/ nitrogen-doped graphene (NG) composite as a catalyst for
Li-O2 batteries, which improved electronic conductivity and promoted reoxidation
reduction reaction (ORR) and oxygen evolution reaction (OER)[39]. In addition, Li et
al. investigated the binary oxide Co2MnO4 as a potential OER catalyst and found that
it can improve stability, maintain good catalytic activity, promote the development of
non-precious metal electrolysis[40]. According to previous reports, due to the
proximity of oxygen and sulfur in the periodic table, SRR catalytic activity has a certain
correlation with ORR catalytic activity. Therefore, it inspired us that the inverse spinel
Co-Mn bimetallic oxides may be able to have an important part in Li-S batteries.

In this study, we developed an inverse spinel Co-Mn bimetallic oxide (i-CMO) as


an improved sulfur reduction reaction (SRR) catalyst. The inverse spinel Co-Mn
bimetallic oxide (i-CMO) has a polar affinity that allows it to capture LiPSs. Besides,
i-CMO is a mixed-valence metal oxide, in which both Co and Mn exist as redox pairs
of +3/+2 valence, which will be conductive to accelerating the reaction kinetics of the
3
conversion of sulfur species. Additionally, when i-CMO is loaded onto a carbon cloth
(CC) substrate to form a flexible free-standing sulfur cathode, such a cathode shows a
satisfactory performance even with high sulfur areal loading.

2. Experiment part

2.1 Material Preparation.

2.1.1 Preparation of Carbon Cloth.

A commercially available carbon cloth (CC, SCC130) was cut into small pieces
of size (2 cm × 3 cm) and hydrophilized through using a mixed solution of nitric and
sulfuric acids. First, 10 wt% nitric acid and 10 wt% sulfuric acid were mixed in the ratio
of 3:1 by volume, and the cut CC were soaked in the mixture and ultrasonicated for 1
h. The carbon cloths were then removed, and the CC were rinsed alternatively with
deionized water (DW) and anhydrous ethanol before being put in a 60 °C oven for 2
hours. Finally, the carbon cloths were removed and set aside.

2.1.2 Preparation of Co2MnO4, CoMn2O4 and Mn2O3.

At room temperature, Mn(NO3)2·4H2O and Co(NO3)2·6H2O in a ratio of 1:2 or


2:1 (total 4.2 mmol) and 16.8 mmol of urea CO(NH2)2 were dissolved in 70 mL of DW
to form a clarified pink solution. The aforementioned solution was put to a 100 mL
autoclave lined with polytetrafluoroethylene and reacted at 100 °C for 6 hours.
Thereafter, the mixed solution was pump-filtered and the resulting sample was placed
in a blower drying oven at 60 ℃ for 3 h. Finally, the resulting material was annealed in
air at 500 °C for 4 h to obtain the normal spinel phase Co2MnO4 or inverse spinel phase
CoMn2O4 composite. As a comparison, Mn2O3 composites were prepared with 4.2
mmol of Mn(NO3)2·4H2O and 16.2 mmol of urea CO(NH2)2 under the same conditions.

2.1.3 Preparation of CC@Co2MnO4, CC@CoMn2O4 and CC@Mn2O3.

In the above steps, a piece of CC (2 cm × 3 cm) was placed among the above
solution and sonicated for 30 min. The solution was transferred to a 100 mL
polytetrafluoroethylene-lined autoclave and reacted at 100 °C for 6 h. Thereafter, the
precursor obtained from the reaction was cleaned with DW and anhydrous ethanol to
remove surface contaminants. Finally, the materials were annealed in air at 400 °C for
4 h to obtain CC@Co2MnO4 and CC@CoMn2O4 composites. As a comparison,
CC@Mn2O3 composites were prepared with 4.2 mmol of Mn(NO3)2·4H2O and 16.2
mmol of urea under the same conditions.

2.1.4 Preparation of Co2MnO4/S, CoMn2O4/S and Mn2O3/S.

Typically, Co2MnO4, CoMn2O4 or Mn2O3, sulfur powder, and C (super P) were


mixed and milled for 10 min in a ratio of 0.07:0.3:0.03 to obtain the mixture. The
4
mixture was then transferred to a vacuum-treated heat-sealed glass tube and heated at
155 ℃ for 12 h to obtain a final homogeneous mixture of sulfur and
Co2MnO4/CoMn2O4/Mn2O3. The final product was placed in a mortar and pestle where
it was ground and named as Co2MnO4/S, CoMn2O4/S and Mn2O3/S composite.

2.1.5 Preparation of CC@Co2MnO4/S, CC@CoMn2O4/S and CC@Mn2O3/S.

The as-prepared CC@Co2MnO4/CC@CoMn2O4/CC@Mn2O3 was immersed in a


CS2 solution containing elemental S for 5 min. To remove excess CS2, the aforesaid
components were removed and dried in a blower drying oven set to 60 °C. Subsequently,
the composites were transferred to a sealed container and heated at 155 °C for 12 h so
that S could be further homogeneously dispersed in the precursors. The obtained final
products were named CC@Co2MnO4/S, CC@CoMn2O4/S and CC@Mn2O3/S.

2.2 Material characterization.

Scanning electron microscopy (SEM, TESCAN MIRA3) is used to observe the


surface topography and dimensions of the material. X-ray diffraction (XRD, D8
Advance, Bruker) is used to determine the composition of the material by comparing it
with a standard card, and X-ray photoelectron spectroscopy (XPS, Thermo Fisher
Scientific K-Alpha, America) records photoelectron energy information to obtain
information about the elements contained on the surface of the material, valence and
chemical shifts. Displacement of the material surface by recording photoelectron
energy information.

2.3 Electrochemical measurement.

2.3.1 Li2S deposition experiment.

Firstly, the electrochemical test parameters of the prepared Li-S battery are
discharged to 2.06 V at constant voltage oof 2.05 V until the current is 0.01 mA. The
current peak will appear after the deposition process is completed. Then, two parts of
Li2S8 reduction and Li2S6 reduction are processed by nonlinear fitting. Finally, the area
of Li2S8 and Li2S6 reduction area is calculated, and the area of deposition area is
calculated by total area.

2.3.2 In situ ultraviolet-visible absorption spectroscopy experiment.

For the in situ Uv-vis spectroscopy experiment, the prepared material is firstly
coated on a clean nickel net to ensure a consistent load for use as a cathode. The lithium
sheet is then cut to specified size and used as the anode. Finally, the in-situ Uv-vis
spectrum battery can be obtained and the measurement is carried out by a workstation
and Uv-vis spectrophotometer.

5
Figure 1. Diagram of an in situ Uv-vis cell.

3. Results and Discussion

To investigate the practical catalytic performance of Co-Mn bimetallic oxides


(CMOs) in Li-S batteries, normal spinel phase and inverse spinel phase CMOs were
synthesized and then loaded on the surface of carbon cloth (CC). As shown in Figure
2, the CC was firstly pretreated with the mixed acids (HNO3 and H2SO4 with volume
ratio of 3:1), and then the CMOs particles were grown in situ on the treated CC via
hydrothermal and high-temperature calcination methods. Depending on the different
ratio of Co and Mn, Co2MnO4 with normal spinel phase (n-CMO) and CoMn2O4 with
inverse spinel phase (i-CMO) can be obtained. Finally, through the melt diffusion
method, S was diffused into CC@CMOs, and the flexible self-supported CC@CMOs/S
cathode was properly manufactured.

Figure 2. Schematic diagram of CMOs synthesis route.

Scanning electron microscopy (SEM) is used to analyze the morphology of the


above samples. Figure 3a-b show SEM images of n-CMO and i-CMO, and it can be
seen that they are the morphology of nanospheres, which may be caused by

6
hydrothermal reaction and reaction conversion between nitrates. In addition, SEM
image of MO in Figure 3c shows regular cube shape with in situ grown on carbon cloth.
Figure 4a-b shows the SEM image of the pristine CC, where a relatively smooth CC
surface can be seen. In contrast, the SEM images of CC@i-CMO (Figure S1) and
CC@MO(n-CMO) (Figure 4c-f) display that some nanoparticles are tightly attached to
the carbon fibers, suggesting that CMOs and MOs were successfully assembled on CC.
In addition, the three-dimensional network with carbon fibers as the backbone provides
a larger surface area and more adsorption active sites, which promotes ion/electronic
diffusion and transport, thereby boosting electrochemical performance[41, 42].

Figure 3. SEM images of (a) n-CMO; (b) i-CMO; (c) MO.

The flexibility and good mechanical properties of CC@CMOs and CC@MO


composites can be demonstrated by folding, rolling and relaxing experiments. As
shown in Figure 4g-i, when the samples are relaxed after folding and rolling
experiments, it can still have good integrity and small deformation, which indicates that
the sample has good flexibility and stability. A certain degree of flexibility is also
beneficial to mitigate the volume expansion effect of the redox kinetics of sulfur[43].

7
Figure 4. SEM images of (a) CC; (c) CC@ n-CMO; (e) CC@MO; High magnification of (b)
CC; (d) CC@n-CMO; (f) CC@MO; The picture of CC: (g) Relaxing, (h) Folding, (i) Rolling.

The as-prepared CMOs and MOs were characterized by XRD to determine their
crystal structures. As shown in Figure 5a, all the main diffraction peaks were located at
2 θ = 18.9°, 31.2°, 36.8°, 44.8°, 55.9° and 65.3°, which can be indexed to the standard
card of n-CMO (PDF#32-0297)[44]. In Figure 5b, all the main diffraction peaks were
situated at 2 θ = 18.2°, 29.3°, 32.9°, 36.3°, 44.8°, 54.4°, 59° and 60.7°, which can be
assigned to the standard card of i-CMO (PDF#18-0408)[45]. Figure 5c shows the XRD
spectrum of MO, and the diffraction peaks were situated at 2 θ = 23.1°, 32.9°, 38.2°,
55.2° and 65.8°, corresponding to the crystal planes of (211), (222), (400), (440) and
(622), respectively (PDF#41-1442)[46]. In addition, the crystal structure shown in the
illustration in the XRD pattern is obtained by the cell parameters on the standard card.
It can be seen that Co is distributed in tetrahedral voids, while Mn is distributed in
octahedral voids, which is called normal spinel and corresponds to Co2MnO4.
Conversely, Co is located in the octahedral void, while Mn is half distributed in the
tetrahedral void, generally distributed in the octahedral void, called inverse spinel,
corresponding to CoMn2O4. The presence of diverse effects in different octahedral
metal atoms leads to the enhanced charge polarization between them, facilitating the
adsorption and deionization of O2. In the inverse spinel phase i-CMO, the charge
polarization effect contributes to the creation of more active adsorption sites, thereby
8
improving catalytic activity and enhancing the kinetics of sulfur reduction[38].

The porous properties of the as-prepared n-CMO, i-CMO and MO materials were
further tested and analyzed by N2 adsorption-desorption isothermal curves. As shown
in Figure 5d-f, the type IV isotherms and typical hysteresis loops indicate the
microporous and mesoporous properties of the three composites. The specific surface
areas of n-CMO, i-CMO and MO materials are 38.69, 47.47, and 16.81 m2 g-1,
respectively. In addition, it can be seen that the adsorption quantity of N2 grows slightly
in the low relative pressure stage, demonstrating the presence of a small number of
micropore pores. Furthermore, the H1-type hysteresis rings also exhibit the
corresponding mesoporous characteristics of the three nano materials. In the same time,
the pore size distribution curves based on Barrett-Joyner-Hallenda (BJH) method are
shown in Figure 5g-i, and it can be seen that the average pore diameters of the three
materials are distributed at 9.01, 10.21 and 10.70 nm, respectively, signifying a typical
mesoporous structure. In general, the large specific surface area and the presence of
micropores and mesopore can provide more adsorption sites and facilitate rapid
ion/electron transport of LiPSs, thus facilitating the rapid conversion reaction of LiPSs
intermediates[47].

Figure 5. XRD pattern of (a) n-CMO, (b) i-CMO, (c) MO; N2 adsorption and desorption
curves of (d) n-CMO, (e) i-CMO, (f) MO; Pore distribution curve of (g) n-CMO, (h) i-CMO;

9
(i) MO

The X-ray photoelectron spectroscopy (XPS) is used as qualitative and semi-


quantitative examination of the chemical states of n-CMO, i-CMO, and MO materials.
In the XPS full-scan spectrum of n-CMO (Figure 6a), the binding energy peaks of the
distinct peaks are located at 285, 531, 642, and 781 eV, which basically correspond to
C 1s, O 1s, Mn 2p, and Co 2p. In the spectrum of C 1s (Figure S2), it is decomposed
into three binding energy peaks at 284.7, 285.5 and 288.7 eV, which are related to C-C
bond, C-O bond and O-C=O bond[48], respectively. In the spectrum of Co 2p, two pairs
of binding energy peaks were shown at 780.2/795.6 and 781.7/797.4 eV, corresponding
to the 2p3/2 and 2p1/2 peaks of Co3+ and Co2+ ions[49], respectively (Figure 6b).
Similarly, in the spectrum of Mn 2p, there are two pairs of binding energy peaks at
641.8/653.4 and 644/654.5 eV, which are related to the spin-orbit splitting of the 2p3/2
and 2p1/2 peaks of Mn3+ and Mn2+ ions[50], respectively (Figure 6c). In addition, a pair
of satellite peaks existed at 785.9/804.3 eV, indicating that Co exists in the n-CMO
composite in the form of +3 and +2 valence states[51]. Among them, Mn and Co with
multiple valence states can provide more polar active sites for anchoring LiPSs[52]. In
the spectrum of O 1s, there are three binding energy peaks at 530.28, 531.78 and 533.28
eV, corresponding to lattice oxygen bonding with metal (Co, Mn) (Ometal-oxygen), non-
lattice oxygen (Onlat) and surface adsorbed oxygen (Oabs)[53, 54], respectively. The
presence of higher levels of non-lattice oxygen and surface-adsorbed oxygen serves
two purposes. Firstly, the stable structures can contribute to the increased electrical
conductivity, thereby promoting the sulfur reduction reaction. Secondly, they provide
additional active sites that effectively adsorb LiPSs intermediates, thereby mitigating
the occurrence of the shuttle effect. As shown in Table S1, the proportion of non-lattice
oxygen and surface adsorbed oxygen in the O1s of i-CMO is much higher than that of
the other two materials. It can be seen that there is a high proportion of surface adsorbed
oxygen in i-CMO, which is more conducive to the redox kinetics of LiPSs.

10
Figure 6. (a) XPS survey of n-CMO, i-CMO and MO; (b) XPS spectra of Co 2p of n-CMO
and i-CMO; (c, d) XPS spectra of Mn 2p and O 1s spectra of n-CMO, i-CMO and MO,
respectively.

Static adsorption studies were performed to validate the adsorption ability of


CMOs and MO on LiPSs. In Figure 7a, the original Li2S6 solution was orange red in
color, and the solution became colorless and transparent after adding CMOs and MO
samples for 24 h. However, the bottle with MO composite showed a yellowish color
relative to the CMOs composite. Besides, the adsorption capacity of CMOs was verified
by Uv-vis absorption spectra. Figure 7b shows that i-CMO sample exhibits the smallest
peak density after the addition of n-CMO, i-CMO and MO, indicating that i-CMO has
stronger anchoring capability and adsorption effects for LiPSs.

In order to further study the electrocatalytic performance of n-CMO, i-CMO and


MO materials, CV tests of symmetric cells based on these three electrodes material
were performed [55]. As demonstrated in Figure 7c, i-CMO exhibits a stronger current
response and a pair of typical redox peaks (located at 0.243 and -0.245V). These results
indicate that i-CMO can promote the redox reaction kinetics of the conversion of
soluble LiPSs.

Li2S deposition experiments were carried out to evaluate the catalytic performance
of the three catalysts. According to Faraday's law and corresponding literature, the Li2S
depositions on different electrodes are further processed[56]. In lithium-sulfur batteries,

11
the deposition process of Li2S involves an electrochemical reaction that occurs during
charging and discharging[57]. During the charging process, lithium ions move from the
negative electrode to the positive electrode, and contact with sulfur (S) on the positive
electrode to form a series of polysulfides, including Li2S4, Li2S6, Li2S8, and so on.
During the discharge process, lithium ions migrate from the positive electrode to the
negative electrode, resulting in the dissolution of a series of polysulfides and the release
of sulfur ions (S2-) and lithium ions (Li+). The released sulfur ion (S2-) returns to the
positive electrode to achieve a charge-discharge cycle[58]. The intensity of the peak
currents and the time of peak appearance are related to the reaction kinetics of Li2S
nucleation, and the integral area represents the amount of the nucleated Li2S[56, 59]. In
Figure 7d-f, the nucleation amount of i-CMO (505.86 mAh g-1) is much larger than n-
CMO (491.58 mAh g-1) and MO (335.01 mAh g-1), indicating that inverse spinel phase
i-CMO can promote polysulfide transformation compared with normal spinel phase n-
CMO and MO. As a result, it can also be seen that bimetallic oxides are more conducive
to promoting the deposition of Li2S and accelerating the kinetics of its transformation
reaction than mono-metallic oxides.

Figure 7g show the EIS of n-CMO at different temperatures. The Nyquist curve is
made up of a semicircle and a straight line that point to the ohmic impedance (Re) and
charge transfer impedance (Rct) [60], respectively. In Figure 7g, it can be seen that the
impedance of the n-CMO material continuously decreases as the temperature increases.
The increase of temperature can usually increase the diffusion rate of lithium ions in
the electrolyte, speed up the chemical reaction rate on the electrode surface, and
accelerate the dissolution and precipitation efficiency of surface sulfide, and thus
resulting in a decrease oof the impedance[57, 58]. Figure S3a shows the EIS of different
materials at 25 ℃. In Table S2, the Rct of i-CMO is lower than the n-CMO and MO,
indicating a quick electron transfer mechanism and reaction kinetics. The EIS data are
further analyzed according to Arrhenius empirical formula. Figure 7h shows that the
logarithm of the rate constant ln(k) has a connection with the reciprocal of the
thermodynamic temperature 1/T and the reciprocal of the charge transfer impedance
Rct, as indicated by Equation (1-2).

Ea
ln 𝑘 = ln 𝐴 ― RT (1)

1
𝑘 ∝ 𝑅𝑐𝑡 (2)

where k is the rate constant, A refers to the pre-factor, Ea is the apparent activation
energy, R is the molar gas constant, and T is the thermodynamic temperature.
Simultaneously, the apparent activation energy Ea of several materials can be
calculated, the results are shown in Figure 7i. Usually, the apparent activation energy
can reflect the speed of the chemical reactions. The apparent activation energy of CMOs

12
cathode is lower than that of MO cathode, and i-CMO has the lowest apparent activation
energy. Therefore, i-CMO shows better catalytic performance than others, thereby it
can better promote the redox kinetics of LiPSs and accelerate the conversion reaction
of active sulfur.

Figure 7. (a)Associated color changes of the Li2S6 solution exposure to n-CMO, i-CMO and
MO; (b) UV-vis absorption spectra for n-CMO, i-CMO and MO; (c) the CV curves of
symmetric cells for n-CMO, i-CMO and MO; Potentiostatic discharge curves of (d) n-CMO,
(e) i-CMO and (f) MO electrode-based lithium polysulfides cells with Li2S8 catholyte;
Impedance at different temperatures of (g) n-CMO; (h) Plot of the logarithm of the rate
constant ln(k) versus the reciprocal 1/T of the thermodynamic temperature; (i) Comparison of
apparent activation energy Ea of different materials.

The transformation process and the concentration of LiPSs were further studied
by via situ Uv-Vis spectroscopy. The redox reaction mechanism of sulfur cathode is
generally accepted as S8→Li2S8→Li2S6→Li2S4→Li2S2/Li2S. First, solid S8 undergoes
a reduction reaction to form a succession of LiPSs such as Li2S8, Li2S6, Li2S4, and so
on. Along with the reaction progresses, these high-order LiPSs degrade into solid
Li2S/Li2S2 [61]. However, during the reduction process, the 𝑆2―
6 ion will undergo a
2― ·― 2― 2― ·―
process of 𝑆6 →𝑆3 →𝑆3 →𝑆4 . Among them, 𝑆3 radical plays an important role

13
in the conversion reaction of LiPSs. With the participation of 𝑆·― 3 radical, the high-
order LiPSs can be successfully converted into low-order LiPSs, which can reduce the
dissolution of LiPSs and inhibit the shuttle effect of LiPSs. Figure 8a-c show that one
of the biggest differences between the patterns of CMOs/S and MO/S cathodes is the
concentration of 𝑆·― ·―
3 radicals located at 619 nm [62]. The concentration of 𝑆3 free
radicals in the cell with CMO as cathodes is higher than that with the MO as cathode.
Especially, in the discharge-charge process of i-CMO/S cathode, the 𝑆·― 3 free radicals
gradually increases in the discharge process and forms an obvious high concentration
center, indicating that the CMO catalyst can effectively promote the conversion process
of 𝑆2―
6 to 𝑆·―
3 , so as to better accelerate the reduction process of active sulfur [63].
Furthermore, for the same current density (0.02 C), the charge and discharge times of
the i-CMO/S cathode (Figure 8d-f) are much longer than those of the n-CMO/S and
MO/S cathodes. suggesting that the i-CMO catalyst achieves a higher utilization rate of
active sulfur. Apparently, above results indicate a good electrochemical performance
of Li-S batteries with i-CMO catalyst.

14
Figure 8. In situ UV–vis spectroscopy of cell with (a) n-CMO, (b) i-CMO and (c) MO;
Related constant current charge/discharge curves of (d) n-CMO, (e) i-CMO and (f) MO.

Figure 9a shows the CV curves for CMOs/S and MO/S cathodes. It can be seen
that the i-CMO/S cathode shows the largest peak current among the three cathodes,
suggesting that it has a good reaction kinetics and low polarization, which is more
favorable to the transformation kinetics of LiPSs. A variety of studies were carried out
to better understand the electrochemical properties of CMOs and MO composites.
Figure 9b-c shows the initial discharge/charge curves of the three sulfur cathodes. The
i-CMO/S cathode displays a high initial discharge specific capacity (1260 mAh g-1)
throughout the activation, which is superior to the n-CMO/S and MO/S cathodes
(Figure S4). In Figure 9d-e, the capacities of i-CMO/S cathode were 1386 mAh g-1 (0.1
C), 1124 mAh g-1 (0.2 C), 962 mAh g-1 (0.5 C), 843 mAh g-1 (1 C), and 728 mAh g-1
(2 C). In contrast, Figure 9f shows that the discharge capacity of n-CMO/S and MO/S
cathodes is less than the i-CMO/S cathodes. Figure 9g shows the cycling performance
of the three cathodes at 0.5 C. The MO/S cathode has a capacity decay rate of 0.25%
per cycle, but the i-CMO/S cathode has good cycling stability with a lower decay rate
of just 0.17% per cycle. In addition, i-CMO and n-CMO based batteries show the
similar cyclic performances at 0.5 C. It is probably because at 0.5 C current density, the
two materials have sufficient time for ion transport and charge transfer, thus they show
similar cyclic performances. When the current density was raised to 1 C in Figure 9h,
the i-CMO/S cathode also demonstrates a good cycling stability with a capacity decay
rate of merely 0.11% per cycle. The high sulfur load capacity of Li-S batteries is one
of the parameters used to evaluate whether these materials can meet the requirement of
practical application. Therefore, CNT/S were coated on carbon cloth (CC) loaded with
different materials, and the performance at different current densities was tested. Figure
9i shows that when the sulfur loading is increased to 3.54 mg cm-2, the i-CMO/S
cathode performs well. At high loading, the discharge specific capacities of i-CMO/S
cathode were 1204 mAh g-1 (0.05 C), 1034 mAh g-1 (0.1 C), 964 mAh g-1 (0.2 C), and
890 mAh g-1 (0.5 C). In contrast, the discharge specific capacities of the n-CMO/S
cathode and the MO/S cathode at different current densities were much lower than the
i-CMO/S cathode.

Even so, the i-CMO/S cathode has yet not reached the performance parameters
that can be directly applied, which may be related to the lack of suitable structure design
of sulfur cathode. However, under the same conditions, the i-CMO/S cathode has
shown significantly better performance than n-CMO/S and MO/S cathodes, which is
enough to support the superiority of the inverse spinel phase Co-Mn bi-metal oxides in
Li-S batteries.

15
Figure 9. (a) Comparison of CV curves at a scanning speed of 0.1mV s-1; Activated constant
current charge/discharge curves of (b) n-CMO and (c) i-CMO; Constant current
charge/discharge curves at different multiplication rates for (d) n-CMO and (e) i-CMO; (f)
Rate performance; (g, h) Cycling performance at 0.5 C and 1 C for n-CMO, i-CMO and MO;
(i) Rate performance under high sulfur loading.

16
Table. 1 Comparison of performances of Li-S batteries with spinel type materials

Sulfur
Reversibl
Rate Cycle Decay loading
e
Material structure performanc numbe rate per mass/ Ref.
capacity/
e r cycle (mg·cm
(mAh·g-1) -2)

NiFe2O4/S 0.1 C 200 963.6 0.15% 1.0~1.3 [64]

0.183
ZnCo2O4@Ti3C2/S 0.5 C 400 1142 1.0~1.5 [65]
%

NiCo2S4/S 1C 300 655 0.13% 1.2 [66]

MnCo2O4@NCFs/ 0.122
0.2 C 300 1177 5 [48]
S %

0.177
CoFe2O4/S 1C 100 1044.2 1 [67]
%

Li2CoTi3O8/S 0.5 C 100 1048 0.3% - [68]

This
CoMn2O4/S 1C 400 949 0.11% 2 wor
k

4. Conclusion

An inverse spinel structure bimetallic oxide i-CMO are prepared and applied in
Li-S batteries, which shows strong interaction with LiPSs and good electrochemical
properties. The CC@i-CMO/S cathode features a three-dimensional network structure
that promotes ion/electron transport and diffusion by providing a larger specific surface
area and more adsorption active sites. In addition, the different valence states of

17
manganese and cobalt in the i-CMO can add more polar active sites, which is more
conducive to promoting the conversion reaction kinetics of LiPSs. Consequently, the
CC@i-CMO/S cathode shows better electrochemical performance, with an initial
discharge specific capacity of 949 mAh g-1 at a high current density of 1 C. After 400
cycles, the discharge capacity is 498 mAh g-1, with a capacity decay rate merely 0.11%
per cycle. Therefore, the use of inverse spinel bimetallic oxide CoMn2O4 (i-CMO) can
accelerate the transformation kinetics of sulfur-containing species, and it shows a good
catalytic ability, which can satisfactorily promote the development of the
industrialization of Li-S batteries.

ASSOCIATED CONTENT

Supporting Information
Electrochemical measurements, SEM images of CoMn2O4 (i-CMO) loaded on
carbon cloth (CC), XPS spectra of C1s of different materials, EIS of CoMn2O4 (i-CMO)
and Mn2O3 (MO) at different temperatures, Re and Rct tables of three materials at 25 ℃,
charge and discharge curves of Mn2O3 (MO).

Acknowledgments
This work was supported by the National Natural Science Foundation of China
(No. 22109135 and No. 52172242), Key Project of Strategic New Industry of Hunan
Province (No. 2019GK2032), Hunan Provincial Natural Scientific Foundation of China
(No. 2022JJ40423), China Postdoctoral Science Foundation (No. 2022TQ0265).

Conflicts of Interest
The authors declare no conflicts of interest.

ORCID:
Xianyou Wang, 0000-0001-8888-6405

REFERENCES

[1] M. Li, J. Lu, Z. Chen, K. Amine, 30 Years of Lithium-Ion Batteries, Adv Mater (2018) 1800561.

[2] J. Zhang, Y. Wu, Y. Xing, Y. Li, T. Li, B. Ren, A review of cathode for lithium-sulfur batteries:
progress and prospects, J Porous Mater (2023).

[3] Q. Shao, S. Zhu, J. Chen, A review on lithium-sulfur batteries: Challenge, development, and
perspective, Nano Research 16 (2023) 8097-8138.

18
[4] R. Mori, Cathode materials for lithium-sulfur battery: a review, J Solid State Electr 27 (2023) 813-
839.

[5] Y. Li, S. Yao, C. Zhang, Y. He, Y. Wang, Y. Liang, X. Shen, T. Li, S. Qin, W. Wen, Molybdenum
carbide nanocrystals modified carbon nanofibers as electrocatalyst for enhancing polysulfides redox
reactions in lithium-sulfur batteries, Int J Energ Res 44 (2020) 8388-8398.

[6] D. Liu, C. Zhang, G. Zhou, W. Lv, G. Ling, L. Zhi, Q.H. Yang, Catalytic Effects in Lithium-Sulfur
Batteries: Promoted Sulfur Transformation and Reduced Shuttle Effect, Adv Sci (Weinh) 5 (2018)
1700270.

[7] J. Yan, X. Liu, B. Li, Capacity Fade Analysis of Sulfur Cathodes in Lithium-Sulfur Batteries, Adv
Sci (Weinh) 3 (2016) 1600101.

[8] C.V. Lopez, C.P. Maladeniya, R.C. Smith, Lithium-Sulfur Batteries: Advances and Trends,
Electrochem 1 (2020) 226-259.

[9] Y. Luo, D. Zhang, Y. He, W. Zhang, S. Liu, K. Zhu, L. Huang, Y. Yang, G. Wang, R. Yu, H. Shu,
X. Wang, M. Chen, Intergrated morphology engineering and alloying strategy for FeNi@NC
Catalysts: Tackling the polysulfide shuttle in Li-S batteries, Chem Eng J 474 (2023).

[10] D. Wang, K. Wang, H. Wu, Y. Luo, L. Sun, Y. Zhao, J. Wang, L. Jia, K. Jiang, Q. Li, S. Fan, J.
Wang, CO2 oxidation of carbon nanotubes for lithium-sulfur batteries with improved electrochemical
performance, Carbon 132 (2018) 370-379.

[11] J. Liu, L. Yi, X. Chen, Y. Tang, Z. Zang, C. Zou, P. Zeng, D. Li, J. Xia, S. Ni, X. Wang, Construction
and Interfacial Modification of a β-PbSnF4 Electrolyte with High Intrinsic Ionic Conductivity for a
Room-Temperature Fluoride-Ion Battery, Acs Appl Mater Inter 15 (2023) 36373-36383.

[12] O. Eksik, S.F. Bartolucci, T. Gupta, H. Fard, T. Borca-Tasciuc, N. Koratkar, A novel approach to
enhance the thermal conductivity of epoxy nanocomposites using graphene core–shell additives,
Carbon 101 (2016) 239-244.

[13] J.-G. Wang, K. Xie, B. Wei, Advanced engineering of nanostructured carbons for lithium–sulfur
batteries, Nano Energy 15 (2015) 413-444.

[14] Z. Zhang, L.L. Kong, S. Liu, G.R. Li, X.P. Gao, A High‐Efficiency Sulfur/Carbon Composite
Based on 3D Graphene Nanosheet@Carbon Nanotube Matrix as Cathode for Lithium–Sulfur Battery,
Adv Energy Mater 7 (2017) 1602543-1602555.

[15] J. Song, M.L. Gordin, T. Xu, S. Chen, Z. Yu, H. Sohn, J. Lu, Y. Ren, Y. Duan, D. Wang, Strong
lithium polysulfide chemisorption on electroactive sites of nitrogen-doped carbon composites for
high-performance lithium-sulfur battery cathodes, Angew Chem Int Edit 54 (2015) 4325-4329.

[16] J. Wang, Y.S. He, J. Yang, Sulfur-based composite cathode materials for high-energy rechargeable
lithium batteries, Adv Mater 27 (2015) 569-575.
19
[17] W. Sun, Y.-C. Lu, Y. Huang, An effective sulfur conversion catalyst based on MnCo2O4.5 modified
graphitized carbon nitride nanosheets for high-performance Li–S batteries, J Mater Chem A 9 (2021)
21184-21196.

[18] X. Liang, C. Hart, Q. Pang, A. Garsuch, T. Weiss, L.F. Nazar, A highly efficient polysulfide
mediator for lithium–sulfur batteries, Nature Communications 6 (2015) 5682-5690.

[19] Y. Miao, Y. Zheng, F. Tao, Z. Chen, Y. Xiong, F. Ren, Y. Liu, Synthesis and application of single-
atom catalysts in sulfur cathode for high-performance lithium–sulfur batteries, Chinese Chem Lett
34 (2023) 107121.

[20] S. Yao, Y. Wang, Y. He, A. Majeed, Y. Liang, X. Shen, T. Li, S. Qin, W. Wen, Synergistic effect
of titanium ‐ oxide integrated with graphitic nitride hybrid for enhanced electrochemical
performance in lithium‐sulfur batteries, Int J Energ Res 44 (2020) 10937-10945.

[21] S. Yao, C. Zhang, F. Xie, S. Xue, K. Gao, R. Guo, X. Shen, T. Li, S. Qin, Hybrid Membrane with
SnS2 Nanoplates Decorated Nitrogen-Doped Carbon Nanofibers as Binder-Free Electrodes with
Ultrahigh Sulfur Loading for Lithium Sulfur Batteries, Acs Sustain Chem Eng 8 (2020) 2707-2715.

[22] J.-L. Qin, H. Zhao, J.-Q. Huang, A metal nitride interlayer for long life lithium sulfur batteries, J
Energy Chem 29 (2019) 1-2.

[23] H. Al Salem, G. Babu, C. V. Rao, L.M.R. Arava, Electrocatalytic Polysulfide Traps for Controlling
Redox Shuttle Process of Li–S Batteries, J Am Chem Soc 137 (2015) 11542-11545.

[24] D. Zhang, Y. Luo, J. Liu, Y. Dong, C. Xiang, C. Zhao, H. Shu, J. Hou, X. Wang, M. Chen, ZnFe2O4–
Ni5P4 Mott–Schottky Heterojunctions to Promote Kinetics for Advanced Li–S Batteries, Acs Appl
Mater Inter 14 (2022) 23546-23557.

[25] F.-J. Liu, W.-L. Luo, Z. Zhang, J. Yu, J.-X. Cai, Z.-Y. Yang, Cation-doped V2O5 microsphere as a
bidirectional catalyst to activate sulfur redox reactions for lithium-sulfur batteries, Chem Eng J 456
(2023) 140948-140955.

[26] J. Feng, C. Shi, H. Dong, C. Zhang, W. Liu, Y. Liu, T. Wang, X. Zhao, S. Chen, J. Song, Design of
ZnSe-CoSe heterostructure decorated in hollow N-doped carbon nanocage with generous adsorption
and catalysis sites for the reversibly fast kinetics of polysulfide conversion, J Energy Chem 86 (2023)
135-145.

[27] Y. He, S. Yao, M. Bi, H. Yu, A. Majeed, X. Shen, Fabrication of ultrafine ZnFe2O4 nanoparticles
decorated on nitrogen doped carbon nanofibers composite for efficient adsorption/electrocatalysis
effect of lithium-sulfur batteries, Electrochim Acta 394 (2021) 139126-139135.

[28] D. Zhang, T. Duan, Y. Luo, S. Liu, W. Zhang, Y. He, K. Zhu, L. Huang, Y. Yang, R. Yu, X. Yang,
H. Shu, Y. Pei, X. Wang, M. Chen, Oxygen Defect ‐ Rich WO3−x–W3N4 Mott–Schottky
Heterojunctions Enabling Bidirectional Catalysis for Sulfur Cathode, Adv Funct Mater 33 (2023).

20
[29] M. Bi, S. Yao, C. Zhang, H. Yu, X. Zhang, H. Liu, T. Zhang, X. Shen, Hybrid of spinel zinc-cobalt
oxide nanospheres combined with nitrogen-containing carbon nanofibers as advanced electrocatalyst
for redox reaction in lithium/polysulfides batteries, Adv Powder Technol 33 (2022) 103710-103720.

[30] J. Habjanic, M. Juric, J. Popovic, K. Molcanov, D. Pajic, A 3D oxalate-based network as a precursor


for the CoMn2O4 spinel: synthesis and structural and magnetic studies, Inorg Chem 53 (2014) 9633-
9643.

[31] D. Meng, Q. Xu, Y. Jiao, Y. Guo, Y. Guo, L. Wang, G. Lu, W. Zhan, Spinel structured CoaMnbOx
mixed oxide catalyst for the selective catalytic reduction of NOx with NH3, Appl Catal B-Environ
221 (2018) 652-663.

[32] Q. Song, Z.J. Zhang, Controlled synthesis and magnetic properties of bimagnetic spinel ferrite
CoFe2O4 and MnFe2O4 nanocrystals with core-shell architecture, J Am Chem Soc 134 (2012) 10182-
10190.

[33] N. Chasserio, B. Durand, S. Guillemet, A. Rousset, Mixed manganese spinel oxides: optical
properties in the infrared range, J Mater Sci 42 (2007) 794-800.

[34] S. Chen, J. Zhang, Z. Wang, L. Nie, X. Hu, Y. Yu, W. Liu, Electrocatalytic NiCo2O4 Nanofiber
Arrays on Carbon Cloth for Flexible and High-Loading Lithium-Sulfur Batteries, Nano Lett 21 (2021)
5285-5292.

[35] J.-X. Lin, X.-M. Qu, X.-H. Wu, J. Peng, S.-Y. Zhou, J.-T. Li, Y. Zhou, Y.-X. Mo, M.-J. Ding, L.
Huang, S.-G. Sun, NiCo2O4/CNF Separator Modifiers for Trapping and Catalyzing Polysulfides for
High-Performance Lithium–Sulfur Batteries with High Sulfur Loadings and Lean Electrolytes, Acs
Sustain Chem Eng 9 (2021) 1804-1813.

[36] Z. Zhang, S. Basu, P. Zhu, H. Zhang, A. Shao, N. Koratkar, Z. Yang, Highly sulfiphilic Ni-Fe
bimetallic oxide nanoparticles anchored on carbon nanotubes enable effective immobilization and
conversion of polysulfides for stable lithium-sulfur batteries, Carbon 142 (2019) 32-39.

[37] T. Sun, X. Zhao, B. Li, H. Shu, L. Luo, W. Xia, M. Chen, P. Zeng, X. Yang, P. Gao, Y. Pei, X.
Wang, NiMoO4 Nanosheets Anchored on N-S Doped Carbon Clothes with Hierarchical Structure as
a Bidirectional Catalyst toward Accelerating Polysulfides Conversion for Li-S Battery, Advanced
Functional Materials 31 (2021) 2101285.

[38] G. Wu, J. Wang, W. Ding, Y. Nie, L. Li, X. Qi, S. Chen, Z. Wei, A Strategy to Promote the
Electrocatalytic Activity of Spinels for Oxygen Reduction by Structure Reversal, Angew Chem Int
Ed Engl 55 (2016) 1340-1344.

[39] Y. Gong, W. Ding, Z. Li, R. Su, X. Zhang, J. Wang, J. Zhou, Z. Wang, Y. Gao, S. Li, P. Guan, Z.
Wei, C. Sun, Inverse Spinel Cobalt–Iron Oxide and N-Doped Graphene Composite as an Efficient
and Durable Bifuctional Catalyst for Li–O2 Batteries, Acs Catal 8 (2018) 4082-4090.

21
[40] A. Li, S. Kong, C. Guo, H. Ooka, K. Adachi, D. Hashizume, Q. Jiang, H. Han, J. Xiao, R. Nakamura,
Enhancing the stability of cobalt spinel oxide towards sustainable oxygen evolution in acid, Nat Catal
5 (2022) 109-118.

[41] R. Elazari, G. Salitra, A. Garsuch, A. Panchenko, D. Aurbach, Sulfur-impregnated activated carbon


fiber cloth as a binder-free cathode for rechargeable Li-S batteries, Adv Mater 23 (2011) 5641-5644.

[42] J.Y. Song, H.H. Lee, W.G. Hong, Y.S. Huh, Y.S. Lee, H.J. Kim, Y.S. Jun, A Polysulfide-Infiltrated
Carbon Cloth Cathode for High-Performance Flexible Lithium-Sulfur Batteries, Nanomaterials-
Basel 8 (2018) 2-10.

[43] Y. Gao, Q. Guo, Q. Zhang, Y. Cui, Z. Zheng, Fibrous Materials for Flexible Li–S Battery, Adv
Energy Mater 11 (2020) 1-24.

[44] K.N. Jung, S.M. Hwang, M.S. Park, K.J. Kim, J.G. Kim, S.X. Dou, J.H. Kim, J.W. Lee, One-
dimensional manganese-cobalt oxide nanofibres as bi-functional cathode catalysts for rechargeable
metal-air batteries, Sci Rep 5 (2015) 7665.

[45] Z. Sadighi, J. Huang, L. Qin, S. Yao, J. Cui, J.-K. Kim, Positive role of oxygen vacancy in
electrochemical performance of CoMn2O4 cathodes for Li-O2 batteries, J Power Sources 365 (2017)
134-147.

[46] Z. Cai, L. Xu, M. Yan, C. Han, L. He, K.M. Hercule, C. Niu, Z. Yuan, W. Xu, L. Qu, K. Zhao, L.
Mai, Manganese oxide/carbon yolk-shell nanorod anodes for high capacity lithium batteries, Nano
Lett 15 (2015) 738-744.

[47] O. Kalawa, P. Kidkhunthod, N. Chanlek, J. Khajonrit, S. Maensiri, Synthesis and electrochemical


properties of polymer solution prepared MnCo2O4 nanoparticles, Ionics 26 (2019) 457-469.

[48] S. Yao, M. Bi, H. Yu, C. Zhang, X. Zhang, H. Liu, T. Zhang, J. Xiang, X. Shen, Spinel manganese-
cobalt oxide nanospheres anchored on nitrogen-containing carbon nanofibers as a highly efficient
redox electrocatalyst in lithium/polysulfides batteries, Appl Surf Sci 598 (2022) 153787-153798.

[49] H. Zhao, Z. Qu, H. Sun, Rational design of spinel CoMn2O4 with Co-enriched surface as high-
activity catalysts for NH3-SCO reaction, Appl Surf Sci 529 (2020).

[50] M. Wei, H. Kang, C. Wang, G. Guo, Y. Liu, Q. Ma, Facet effect of MnCo2O4 nanocrystals for
enhanced oxygen reduction reaction in alkaline medium, Appl Surf Sci 631 (2023).

[51] L. Zhang, G. He, S. Lei, G. Qi, H. Jiu, J. Wang, Hierarchical hollow microflowers constructed from
mesoporous single crystalline CoMn2O4 nanosheets for high performance anode of lithium ion
battery, J Power Sources 326 (2016) 505-513.

[52] C. Li, X. Han, F. Cheng, Y. Hu, C. Chen, J. Chen, Phase and composition controllable synthesis of
cobalt manganese spinel nanoparticles towards efficient oxygen electrocatalysis, Nat Commun 6
(2015) 7345.
22
[53] B. Pattanayak, F.M. Simanjuntak, D. Panda, C. Yang, A. Kumar, P. Le, K. Wei, T. Tseng, Role of
precursors mixing sequence on the properties of CoMn2O4 cathode materials and their application in
pseudocapacitor, Sci Rep 9 (2019) 16852.

[54] S. Qiao, Q. Wang, D. Lei, X. Shi, Q. Zhang, C. Huang, A. Liu, G. He, F. Zhang, Oxygen vacancy
enabled fabrication of dual-atom Mn/Co catalysts for high-performance lithium–sulfur batteries, J
Mater Chem A 10 (2022) 11702-11711.

[55] X. Wu, N. Liu, B. Guan, Y. Qiu, M. Wang, J. Cheng, D. Tian, L. Fan, N. Zhang, K. Sun, Redox
Mediator: A New Strategy in Designing Cathode for Prompting Redox Process of Li-S Batteries,
Adv Sci (Weinh) 6 (2019) 1900958.

[56] J.-L. Yang, D.-Q. Cai, Q. Lin, X.-Y. Wang, Z.-Q. Fang, L. Huang, Z.-J. Wang, X.-G. Hao, S.-X.
Zhao, J. Li, G.-Z. Cao, W. Lv, Regulating the Li2S deposition by grain boundaries in metal nitrides
for stable lithium-sulfur batteries, Nano Energy 91 (2022) 106669.

[57] Z. Li, Y. Zhou, Y. Wang, Y.C. Lu, Solvent ‐ Mediated Li2S Electrodeposition: A Critical
Manipulator in Lithium–Sulfur Batteries, Adv Energy Mater 9 (2019) 1802207.

[58] F.Y. Fan, W.C. Carter, Y.M. Chiang, Mechanism and Kinetics of Li2S Precipitation in Lithium–
Sulfur Batteries, Adv Mater 27 (2015) 5203-5209.

[59] M. Chen, X. Zhao, Y. Li, P. Zeng, H. Liu, H. Yu, M. Wu, Z. Li, D. Shao, C. Miao, G. Chen, H. Shu,
Y. Pei, X. Wang, Kinetically elevated redox conversion of polysulfides of lithium-sulfur battery
using a separator modified with transition metals coordinated g‑C3N4 with carbon-conjugated, Chem
Eng J 385 (2020) 123905-123917.

[60] X. Li, A. Lushington, Q. Sun, W. Xiao, J. Liu, B. Wang, Y. Ye, K. Nie, Y. Hu, Q. Xiao, R. Li, J.
Guo, T.K. Sham, X. Sun, Safe and Durable High-Temperature Lithium-Sulfur Batteries via
Molecular Layer Deposited Coating, Nano Lett 16 (2016) 3545-3549.

[61] H. Chen, Z. Wu, M. Zheng, T. Liu, C. Yan, J. Lu, S. Zhang, Catalytic materials for lithium-sulfur
batteries: mechanisms, design strategies and future perspective, Mater Today 52 (2022) 364-388.

[62] Z. Ma, Y. Liu, J. Gautam, W. Liu, A.N. Chishti, J. Gu, G. Yang, Z. Wu, J. Xie, M. Chen, L. Ni, G.
Diao, Embedding Cobalt Atom Clusters in CNT-Wired MoS2 Tube-in-Tube Nanostructures with
Enhanced Sulfur Immobilization and Catalyzation for Li-S Batteries, Small 17 (2021) 2102710.

[63] H. Yuan, W. Zhang, J.-g. Wang, G. Zhou, Z. Zhuang, J. Luo, H. Huang, Y. Gan, C. Liang, Y. Xia,
J. Zhang, X. Tao, Facilitation of sulfur evolution reaction by pyridinic nitrogen doped carbon
nanoflakes for highly-stable lithium-sulfur batteries, Energy Storage Mater 10 (2018) 1-9.

[64] Z. Zhang, D.-H. Wu, Z. Zhou, G.-R. Li, S. Liu, X.-P. Gao, Sulfur/nickel ferrite composite as cathode
with high-volumetric-capacity for lithium-sulfur battery, Sci China Mater 62 (2018) 74-86.

[65] A. Wei, L. Wang, Z. Li, Metal-organic framework derived binary-metal oxide/MXene composite as
23
sulfur host for high-performance lithium-sulfur batteries, J Alloy Compd 899 (2022) 163369-163378.

[66] R. Li, Z. Bai, W. Hou, J. Qiao, W. Sun, Y. Bai, Z. Wang, K. Sun, Spinel-type bimetal sulfides
derived from Prussian blue analogues as efficient polysulfides mediators for lithium−sulfur batteries,
Chinese Chem Lett 32 (2021) 4063-4069.

[67] J. Pu, Y. Tan, T. Wang, X. Zhu, S. Fan, Ultrathin Two-Dimensional Fe–Co Bimetallic Oxide
Nanosheets for Separator Modification of Lithium–Sulfur Batteries, Molecules 27 (2022) 7762.

[68] M. Qian, Y. Tang, L. Liu, Y. Gao, X. Li, Well-dispersed Li2CoTi3O8 nanoparticles as a


multifunctional material for lithium-ion batteries and lithium-sulfur batteries, J Alloy Compd 896
(2022) 162926-162934.

24
Highlights

 The Co-Mn bimetallic oxides with spinel structure are successfully synthesized.

 The i-CMO catalyst realizes the excellent capture-adsorption-catalysis to LiPSs.

 The catalysis of i-CMO is better than n-CMO in promoting the conversion of LiPSs.

 The i-CMO-based Li-S batteries exhibit a better performance than that of n-CMO.

25

You might also like