You are on page 1of 13

FULL PAPER

Lithium-Ion Batteries www.afm-journal.de

Tuning Anionic Redox Activity and Reversibility


for a High-Capacity Li-Rich Mn-Based Oxide Cathode via
an Integrated Strategy
Qingyuan Li, Dong Zhou, Lijuan Zhang, De Ning, Zhenhua Chen, Zijian Xu, Rui Gao,
Xinzhi Liu, Donghao Xie, Gerhard Schumacher, and Xiangfeng Liu*

1. Introduction
When fabricating Li-rich layered oxide cathode materials, anionic redox
chemistry plays a critical role in achieving a large specific capacity. Lithium-ion batteries (LIBs) are criti-
Unfortunately, the release of lattice oxygen at the surface impedes the cally important components in portable
reversibility of the anionic redox reaction, which induces a large irreversible artificial intelligence devices, electric
capacity loss, inferior thermal stability, and voltage decay. Therefore, vehicles (EVs), and smart grid storage
systems.[1] More than ever before, safe,
methods for improving the anionic redox constitute a major challenge
stable, and high-performance LIBs are
for the application of high-energy-density Li-rich Mn-based cathode urgently needed, as the rise of global EVs
materials. Herein, to enhance the oxygen redox activity and reversibility in continues, because the specific capacity
Co-free Li-rich Mn-based Li1.2Mn0.6Ni0.2O2 cathode materials by using an of the current LIBs cannot meet the
integrated strategy of Li2SnO3 coating-induced Sn doping and spinel phase demands of the market.[2] In LIB sys-
tems, the weight and cost of the cathode
formation during synchronous lithiation is proposed. As an Li+ conductor,
materials account for ≈40–50% of the
a Li2SnO3 nanocoating layer protects the lattice oxygen from exposure at entire battery, and the specific capacity
the surface, thereby avoiding irreversible oxidation. The synergy of the of the cathode materials is significantly
formed spinel phase and Sn dopant not only improves the anionic redox lower than that of the anode materials
activity, reversibility, and Li+ migration rate but also decreases Li/Ni mixing. used in the current study. Therefore,
The 1% Li2SnO3-coated Li1.2Mn0.6Ni0.2O2 delivers a capacity of more than cathode materials remain the main bot-
tleneck and to overcome this challenge,
300 mAh g−1 with 92% Coulombic efficiency. Moreover, improved thermal
high-energy-density cathode materials
stability and voltage retention are also observed. This synergic strategy may are highly desired.
provide insights for understanding and designing new high-performance Among various cathode mate-
materials with enhanced reversible anionic redox and stabilized surface rials, Li- and Mn-rich layered oxides
lattice oxygen. (LMLO), which can be described as
xLi2MnO3 · (1−x) LiMO2 (M = Ni, Co, Mn,
etc.), have attracted an increasing amount
of attention as the most promising candi-
dates due to their high specific capacity (≈300 mAh g−1) and
high energy density (≈1000 Wh kg−1).[1a,3] In addition to the
Dr. Q. Li, Dr. R. Gao, D. Xie, Prof. X. Liu well-known redox behavior of the transition metal (TM) ions
Center of Materials Science and Optoelectronics Engineering used for charge compensation during Li+ insertion/extraction,
College of Materials Science and Optoelectronic Technology the anionic redox chemistry that occurs in Li-rich layered oxide
University of Chinese Academy of Sciences
Beijing 100049, P. R. China
cathode materials also plays a critical role in the large specific
E-mail: liuxf@ucas.ac.cn capacity, as confirmed by several research groups.[3a,4] Unfor-
Dr. D. Zhou, Dr. D. Ning, Dr. X. Liu, Prof. G. Schumacher tunately, the irreversible release of lattice oxygen at the sur-
Helmholtz-Center Berlin for Materials and Energy face impedes the reversibility of the anionic redox, which can
Hahn-Meitner-Platz 1, Berlin 14109, Germany also induce a large irreversible capacity loss, inferior thermal
Prof. L. Zhang, Prof. Z. Chen, Prof. Z. Xu stability, and even voltage decay.[5] Therefore, approaches
Shanghai Synchrotron Radiation Facility for improving the anionic redox chemistry (O2−→O2n−) while
Shanghai Institute of Applied Physics
Chinese Academy of Sciences suppressing the irreversible O2 release are a big challenge for
Shanghai 201204, P. R. China the continued development of high-energy Li-rich cathode
The ORCID identification number(s) for the author(s) of this article materials.[3a,4e,6]
can be found under https://doi.org/10.1002/adfm.201806706. Although some have reported that the incorporation of
precious 4d (Ru) and 5d (Ir) TM ions can mitigate the O2
DOI: 10.1002/adfm.201806706 gas loss,[4f,7] in contrast, cheap 3d (Ni and Mn) TM ion-based

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (1 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

cathode materials are urgently needed to obtain the high spe- 2. Results and Discussion
cific capacity and excellent cycle performance required for
enhanced LIB performance. In such cases, the ions contribute 2.1. Crystal Structure and Morphology
to these properties by regulating the reversible oxygen redox
chemistry. However, the low-cost 3d TM-based cathode mate- X-ray diffraction (XRD) was used to determine the phase com-
rials, such as Li- and Mn-rich layered oxide cathodes, suffer position of the pristine and Li2SnO3-coated samples (1%, 3%,
from several intrinsic drawbacks, such as a low lithium ion and 5% LSO). As shown in Figure  1a, the main peaks of the
conductivity, structural instability, and corrosion of the TM four samples are similar due to the common hexagonal R3m
elements in the electrolyte.[3b,8] These drawbacks cause an lattice. In particular, the intensities of the (003), (104), and (101)
inferior rate capability and poor cycling performance, signifi- peaks of the pristine materials are stronger than those of the
cant voltage fading, and low Coulombic efficiency. In LMLO 1%, 3%, and 5% LSO, which indicates that the coating layer has
cathodes, the transformation from a layered structure to a been prepared successfully. A few broad superlattice reflections
spinel phase results in a crystal instability when the cathode appear at ≈20–25° with a space group of C2/m and are caused
materials are charged to 4.8 V. Although the high voltage is by the ordering of Mn and Li atoms in the TM layers, indicating
conducive to improving the energy density because of the elec- the presence of Li2MnO3.[13] In addition to the similar diffrac-
trochemical reaction of the Ni2+/3+ and Ni3+/4+ couples that tion peaks of the layered cathode materials, the peaks of the
occurs at ≈4.7 V,[3a,9] nonetheless the voltage also causes sev- spinel phase can be clearly identified in the 1%, 3%, and 5%
eral undesirable side reactions. For example, as the fraction LSO coating samples. The strong cubic spinel diffraction peaks
of material in the spinel phase increases during prolonged of the 1%, 3%, and 5% LSO samples are readily indexed as the
cycling, the capacity and voltage fade due to the dissolution space group of the Fd-3m structure.[9,11f ] The spinel phase in
of bivalent manganese near the 3 V plateau.[8,10] However, the 3% and 5% LSO samples is more obvious than in the 1%
some reports have demonstrated that the electrochemical LSO sample based on the intensity of the diffraction peak at
performance of Li-rich layered cathodes benefits from the ≈44°.[12k,14] In addition, as the content of the Li2SnO3 coating
integration of a tiny amount of spinel phase during mate- layer increases, new peaks appear at 25–35° in the 3% and 5%
rial synthesis.[10c,11] Moreover, to suppress the dissolution of LSO samples, to which we ascribe the tetragonal SnO2 phase
Mn2+ and release of oxygen, several strategies, such as bulk with a space group of P42/mnm.[15] Furthermore, to clearly
doping and surface coating, have been reported.[5,12] However, demonstrate the differences between the XRD patterns of the
with regard to lattice doping, surface modification and spinel pristine, 1%, 3%, and 5% LSO samples, enlarged plots of
integration, the overall contributions of oxygen redox chem- the 003 peaks are shown in Figure 1b. Compared with the
istry continue to receive little attention due to the complicated pristine material, the 003 peak of 1% LSO shifts to a lower
nature of these processes. 2θ angle, which indicates that the 003 slabs have expanded
In this work, we propose a facile strategy that simultane- due to the doping and migration of Sn4+ during the coating
ously combines the advantages of doping, surface coating, and process.[12e,16] This expansion is beneficial to the lithium inser-
spinel phase integration to improve the oxygen redox activity tion/extraction process. However, the peaks of 3% LSO and
and suppress the irreversible loss of lattice oxygen as well as 5% LSO clearly show a shift to higher angle, which indicates
alleviate the other drawbacks of LMLO. A nanolayer of Li+- a decrease of the interplanar spacing of the 003 crystal plane.
conductive Li2SnO3 is coated on the surface of LMLO using a This phenomenon may be caused by the dissolution of Li+ out
synchronous lithiation strategy. During lithiation, Sn ions dif- of the bulk phase from the uneven coating layer during the
fuse into the surface of the LMLO cathode materials in a pillar high-temperature synchronous lithiation reaction,[16b] which
shape, and a small amount of spinel phase forms simultane- is confirmed by elemental mapping images obtained by trans-
ously. In this process, the nano-Li2SnO3 coating layer not only mission electron microscopy (TEM) (see below).
has a higher Li+ conductivity than other conventional metal- Rietveld refinement was performed to further confirm
oxide coatings but also induces Sn doping and the formation the effects of Sn4+ doping in the 1%, 3%, and 5% LSO sam-
of a spinel-layered heterogeneous phase. The synergistic effect ples. The fitting curves and Rietveld refinement data of XRD
of Sn doping, nano-Li2SnO3, and spinel phase integration powder diffraction data are shown in Figure S1 and Table S1
enhances the reversible anionic redox activity and suppresses (Supporting Information). Based on the Rietveld refinement,
the lattice oxygen loss from the surface. Aberration-corrected the contents of the layered and spinel phases in 1% LSO are
scanning transmission electron microscopy (AC-STEM) and ≈79.5% and 20.5%, respectively. As shown in Table S1 of the
neutron diffraction with Rietveld refinement were used to Supporting Information, the lattice parameter c of the pris-
verify the existence of the spinel phase and the reduction tine material is 14.2688 Å, whereas that of the 1% LSO with a
of Li/Ni mixing. Ex situ soft X-ray absorption spectroscopy uniform coating layer increases to 14.2758 Å, suggesting that
(XAS), ex situ X-ray photoelectron spectroscopy (XPS), and Sn has diffused into the bulk phase from the coating layer.
differential scanning calorimetry (DSC) were applied to probe Interestingly, with the increase of the coating content, the lat-
the enhancement of anionic redox reactivity and reversibility. tice parameter c of the 3% and 5% LSO samples gradually
We believe that the proposed synergistic strategy could give decreases to 14.2506 and 14.2444 Å, respectively, which may
new insights into the design of high-performance Li+, Na+, be due to an uneven coating layer. Importantly, this change is
and other types of battery cathodes by regulating the anionic consistent with the results from the XRD analysis. The slab
redox of bulk phase and protecting the lattice oxygen at the thickness (S(MO2)) and the interslab thickness (I(LiO2)) for the
surface. layered phase can be calculated by the following two formulas

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (2 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 1.  XRD and Rietveld refinement patterns from neutron powder diffraction. a) Overall XRD patterns of the pristine, 1% LSO, 3% LSO, and 5%
LSO samples; b) magnified patterns of the four samples between 18 and 19.5°. Rietveld refinement of the neutron powder diffraction patterns for the
c) pristine material and d) 1% LSO sample. Rwp values for the pristine material and 1% LSO sample are 5.62% and 6.14%, respectively.

(S(MO2)  = 2[(1/3) – zox] c; I(LiO2)  = (c/3) − S(MO2)).[16b] According has a lower degree of cation mixing than the pristine sample. It
to this equation, the I(LiO2) values of the pristine material and is noteworthy that the I(003)/I(104) value of 3% LSO is higher
1% LSO are calculated to be 2.1728 and 2.1790 Å, respectively, than that of the pristine material, which could be responsible
which further demonstrates that Sn has diffused into the sur- for its slightly better cycle performance at low rate. From these
face and bulk phase and that doping with Sn4+ has induced results, we can conclude that the uniform Li2SnO3 nanocoating
the spinel phase and expanded the interslab space. Enlarge- induced trace Sn4+ doping in the surface and a small amount
ment of the interslab space can reduce the Li activation energy, of spinel phase formation by the synchronous lithiation route.
which can significantly enhance the rate of Li+ diffusion. The The combined effect of Sn doping and spinel phase formation
S(MO2) values of the pristine sample and 1% LSO are calcu- has increased the interslab spacing, which can reduce the polar-
lated to be 2.5835 and 2.5796 Å, respectively. Compared with ization of the redox reaction because of the lower Li activation
the pristine sample, the decreased S(MO2) value of 1% LSO is energy.
conducive to the stability of the layered structure. However, Although XRD patterns can give important structural infor-
with the increasing coating content, the values of S(MO2) and mation, they are not sensitive to the light Li atom. Because Li+
I(LiO2) change irregularly because of the nonuniform coating (0.72 Å) and Ni+ (0.69 Å) have similar ionic radii, Li/Ni mixing
thickness. Furthermore, the value of I(003)/I(104), which is an is inevitable, but the degree of mixing cannot be quantita-
indicator of Li/Ni cation mixing, is also affected. It is believed tively analyzed by XRD. Instead, neutron diffraction is needed
that the higher I(003)/I(104) value means a lower degree of to quantitatively detect Li occupancy. Figure 1c,d shows the
Li/Ni cation mixing, which is beneficial to the rate capability refinements of the neutron powder diffraction patterns for
and the cyclic capability. According to the XRD results, the the pristine material and 1% LSO sample. The Rwp values of
I(003)/I(104) value of the pristine specimen is 1.7293, which Rietveld refinement of the pristine sample and 1% LSO are
is lower than 2.1333 in 1% LSO. This suggests that 1% LSO 6.14% and 5.62%, respectively. The Rietveld refinements show

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (3 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 2.  Morphological and structural characterization of the 1% LSO sample. a) TEM image of 1% LSO. b) HRTEM image of 1% LSO, from the inset
rectangle in (a). c,d) The SAED pattern and identities of the different diffraction rings, corresponding to the layered, spinel, and Li2SnO3 structures,
respectively. e–i) STEM image and EDS maps of O, Mn, Ni, and Sn.

that the Li/Ni mixing of the pristine and 1% LSO samples are Figure 2a shows a TEM micrograph of the morphology of 1%
5.4% and 1.3%, respectively, with the detailed site occupancy LSO, which was found to exhibit uniform spherically shaped
information presented in Table S2 (Supporting Information). nanoparticles on its surface. These shape observations agree
This result clearly illustrates that an appropriate amount of well with the imaging results obtained with FESEM presented
Li2SnO3 coating can induce Sn doping and spinel phase forma- in Figure S2b (Supporting Information). Figure 2b shows a
tion. These changes can significantly suppress Li/Ni mixing, high-resolution TEM (HRTEM) image captured in the region of
which is beneficial for Li+ diffusion and rate performance. The the red rectangle in Figure 2a. A heterostructure with three dis-
reduced Li/Ni mixing can also be inferred from the following tinct layers can clearly be observed in Figure 2b. The thickness
AC-STEM results. of the outermost layer is ≈2 nm, and the lattice fringe spacing
The morphologies of the four samples were characterized is 0.26 nm, which is attributed to the (−131) planes of Li2SnO3.
by the field emission scanning electron microscopy (FESEM) The interslab spacing of the intermediate layer is 0.24 nm,
(Figure S2, Supporting Information). All samples exhibit spherical whereas the spacings of the innermost layers are 0.47 and
particles with diameters of ≈1–2 μm, with the smaller particles 0.25 nm; the former corresponds to the (222) plane of the spinel
observed at the sample surface. The pristine material particles phase, and the latter to the (003) and (101) planes of the layered
have rough surfaces with a small amount of particle aggregation structure. A selected-area electron diffraction (SAED) image of
(Figure S2a, Supporting Information). Compared with pristine this region (Figure 2c) and the corresponding diffraction pat-
material, the 1% LSO sample exhibits more spherically shaped tern (Figure 2d) further verified this heterostructure. Three
particles and a smoother surface (as shown in Figure S2b of the diffraction rings are clearly visible in the SAED pattern, corre-
Supporting Information). However, the 3% and 5% LSO samples sponding to the (101), (222), and (−131) planes of the layered
shown in Figure S2c,d (Supporting Information) possess rougher structure, spinel phase, and Li2SnO3, respectively. All of these
surfaces with larger particles. As a result, sample morphology results support the interpretation of a high degree of structural
obviously changes with the increasing content of Li2SnO3 coating, compatibility between the layered phase and the spinel struc-
which might result from the uneven coating thickness. ture, due to the close cubic packing of these structures.[11c] To

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (4 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 3.  HAADF-STEM images of the pristine material and 1% LSO sample. a) HAADF-STEM image of the pristine material. Bright spots represent
transition metals (Ni or Mn); the dark spots represent Li layers; the inset rectangle and the histogram represent the mixing of Li/Ni. b) Crystal structure
models of the layered phase with Li/Ni mixing. c) HAADF-STEM image of the 1% LSO sample. The bright dots with arrows illustrate the Sn doped into
the TM layer in the inset color image; the green rectangle represents the spinel structure. d) Crystal model of the spinel phase.

directly observe the distribution of the Li2SnO3 nanocoating on rate property of this material.[12e] Moreover, Figure 3c confirms
the surface of the materials, EDS mapping was also employed the existence of the spinel phase at an atomic level based on
while analyzing the samples using TEM, and the results are observations consistent with the typical atomic arrangement
presented in Figure 2e–i and Figure S3 ii and iii (Supporting of the spinel structure (Figure 3d). Since the spinel structure
Information). Accordingly, all TM elements are uniformly possesses 3D diffusion channels that are available to Li+, this
distributed in the 1% LSO specimen. HRTEM images of the feature also improves the rate performance of the material.
pristine, 3% and 5% LSO samples are also shown in Figure S3 Based on the above HRTEM analyses and XRD results, we can
(Supporting Information). Three phases are demonstrated in conclude that the Li2SnO3 nanocoating has induced the forma-
Figure S3 II and III of the Supporting Information, revealing tion of a spinel phase, as well as a doping effect that expands
a structure similar to that of the 1% LSO specimen. In com- the lattice spacing, decreases the degree of cation mixing and
parison, Sn is not distributed uniformly on the surface of the reduces the activation energy of Li+ migration due to the elec-
3% and 5% LSO samples, which could deteriorate the electro- trostatic interactions.
chemical performance of these materials.
To obtain more detailed information about the degree of Sn
doping and spinel phase formation, AC-STEM with HAADF 2.2. Electronic Structure Analysis
detector was also employed. When using this technique, Mn,
Ni, and Sn atoms appear brighter than Li and O atoms, due To further explore the surface states of the samples, XPS
to the differences in atomic numbers. As shown in Figure  3a, analy­ses of bare and coated samples were performed (as shown
the HAADF-STEM pattern of the pristine sample reveals a in Figure S4 of the Supporting Information). According to
perfectly layered structure. The TM layers are indicated by Figure S4a of the Supporting Information, the existence of
the bright spots, which reside in a straight line (plane), with Sn4+ is proven by the peaks corresponding to Sn 3d3/2 and
the lithium slabs located between these TM layers.[17] Obvious 3d5/2, which are located at 494.8 and 486.4 eV, respectively,
bright spots appear within the Li slabs (dark dots), as identi- and confirm that the Li2SnO3 structure is coated on the sur-
fied with the rectangle, indicating a degree of Li/Ni mixing in face of the materials.[16a] In Figure S4b of the Supporting
the pristine materials.[3b] In addition, clear contrast peaks in a Information, two O 1s peaks can be observed for the pristine
corresponding histogram also illustrates the existence of Li/Ni material. The main peak and its shoulder peak are located at
mixing. Figure 3b shows an atomic model of Li/Ni mixing in 529.3 and 531 eV, respectively. The main peak is attributed
the layered phase. Figure 3c shows an HAADF-STEM image to the lattice oxygen, whereas the shoulder peak is associ-
of 1% LSO. Several bright contrast spots can be observed ated with the absorbed oxygen on the surface of the material,
between two bright spot sites, as marked with white arrows, which corresponds to species of Li2CO3 and LiOH.[4e] Com-
which stand out structurally when compared with the pris- pared with the pristine material, the shoulder peaks of the
tine material (inset in Figure 3c), implying that the Sn atom coated samples have extremely low relative intensities, which
is only doped into the TM layer rather than the Li slabs. This means that very little Li2CO3 and LiOH exist on the surface of
structure promotes Li+ diffusion and leads to the high diffusion the coated material. In addition, the increased binding energy

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (5 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

of O 1s is related to the doping of Sn4+ in the coated mate-


rials, which is caused by the stronger bonding between O
and Sn4+ (△HfSn-O = 548 kJ mol−1>△HfMn-O = 402 kJ mol−1 or
△HfNi-O  = 391.6 kJ mol−1) and the associated higher electron-
egativity of this bond; this feature indicates a higher degree
of protection of the layered oxides in Li-rich materials.[16a]
Figure S4c,d of the Supporting Information shows the Ni 2p
and Mn 2p XPS spectra of the four samples. Compared with
the peaks observed for the bare material, the Ni 2p peaks of the
1% Li2SnO3 nanocoated samples are clearly shifted to a lower
binding energy, which implies that the valence state of the Ni
has been reduced, as well as the covalence of NiO. This result
is attributed to the doping of Sn4+ and the need for electric
neutrality to be maintained in the material,[18] which changes
the local geometry of the electronic structure. The peak of Mn
2p is located at ≈642.3 eV in all four samples, corresponds to
Mn4+, and this does not obviously change, which reflects the
stability of the layered structure.
To further study the relationship between the surface and
bulk material of the four samples, soft XAS was also applied,
allowing for the electronic structures of the TMs to be com-
pared between the bare and coated samples. The L-edge XAS
spectra of Ni and Mn were measured in TEY mode (a probe
depth of ≈5 nm) are shown in Figure 4a,b. The L2 and L3 edges
of Ni in the coated materials shift to lower energy compared
with those of the bare samples. The L2 and L3 edges correspond
to the transitions from the 2p3/2 and 2p1/2 core levels to unoc-
cupied 3d holes.[19] The shift of these peaks to lower energies
implies that both the covalence of the NiO bond and the
valence of Ni have been reduced.[19a] This result can be ascribed
to the effect of Sn doping near the surface, which changes the
local electronic structure due to the stronger covalent bond of
SnO and the requirement to maintain charge balance; fur-
thermore, this result is consistent with the results of XPS.
Figure 4b shows the Mn L2 and L3 edge spectra. There is no
obvious change in the L2 and L3 peaks of Mn, indicating the sta-
bility of the structure due to the Mn4+ valence state. Figure 4c
shows the O-K edge spectra recorded for the four samples in
TEY mode. The features in these spectra can be attributed to
the transition of O 1s to the hybridized state of O 2p and TM 3d
orbitals with peaks below ≈534 eV. Similarly, the broad peaks at
energies greater than 534 eV can be ascribed to the transition
of the O1s electron to the hybridized state of TM 4sp and O
2p orbitals.[19a,b] In addition, the intensity of the O K-edge peak
is strongly related to the covalence of TMO bonding.[19a,20]
The associated peaks of the 3% and 5% LSO samples present
at ≈538 eV are stronger than those of the bare and 1% LSO
samples, indicating the strong covalency between the TM and
O in the 3% and 5% LSO samples; this strong covalency is
detrimental to battery performance because of the evolution
of O2.[19a] Moreover, to verify whether the nanocoating modi-
fied only the surface of the pristine materials and not the bulk
phase, XAS spectra were collected in FY mode (a probe depth
of ≈100 nm), as shown in Figure S5 of the Supporting Infor-
mation. According to this figure, no changes were observed
for the TM and oxygen atoms, which means that the electronic
structure is stable in the bulk phase; this stability is beneficial
to the desired cyclic performance and reversible oxygen redox Figure 4.  XAS Spectra of the four samples in TEY Mode. a) Ni L2,3 edge.
chemistry.[12c,19b] b) Mn L2,3 edge. c) O K edge.

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (6 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

The above analyses clearly illustrates that the Li2SnO3 nano- as observed for all samples.[21] The prominent plateau at ≈4.5 V
coating was deposited successfully onto the surface of the Li- indicates the irreversible removal of Li and O atoms from
rich Li1.2Mn0.6Ni0.2O2 substrate during the lithiation process. In Li2MnO3, and the activation of this layered component.[11a]
addition, the analyses indicate that part of the Sn has diffused The plateau is longer in the pristine material, implying that
into the bulk phase, decreasing the covalence between the TM the activation is more complete in the first charge process. It is
and oxygen species, which can improve the properties of the noteworthy that there are two short plateaus at ≈4.75 V, corre-
battery by alleviating oxygen evolution. Furthermore, the 1% sponding to the Li+ deintercalation from the 8a tetrahedral sites
LSO nanocoating with a modified surface structure is condu- and the oxidation of Ni2+ to Ni4+ in the spinel phase; this pheno­
cive to the electrochemical performance and charge compensa- menon is also apparent during the discharge process in com-
tion via oxygen. parison with the pristine material.[22] The typical discharge curve
characteristics of the spinel structure are shown in Figure 5a
and Figure S6 (Supporting Information). In addition to the
2.3. Electrochemical Performance discharge profile at ≈4.75 V, as discussed above, the voltage
plateaus (≈2.7 and 2.2 V) can be clearly observed in the coated
The initial charge–discharge curves of the pristine and materials but are not present in the pristine sample, indicating
coated materials were tested at a current density of 0.05 C that the nanocoating has successfully induced the formation of
(1 C = 250 mAh g−1) and a voltage of 2–4.8 V, as shown in a spinel phase. These two discharge plateaus can be ascribed to
Figure  5a and Figure S6 (Supporting Information). During a phase transition and the reduction of Mn4+ to Mn3+.[23] The
charging, the voltage plateau below 4.5 V corresponds to Li+ initial discharge capacities of the pristine and nanocoated (1%,
extraction from the layered structure of the sample and the oxi- 3%, and 5% LSO) samples are 247, 244, 125, and 106 mAh g−1,
dation of the extremely few Mn3+ ions in the spinel structure, respectively. The discharge specific capacity gradually decreases

Figure 5.  a) Initial charge–discharge curves at 0.05 C for the pristine sample and 1% LSO; b) The rate performance. c–e) The cyclic performance at 0.2,
1, and 5 C under room temperature, and f) the cyclic performance at 1 C at 55 °C. g) The average discharge voltage profile versus long-term cycling
at 1 C. h) DSC profiles of the pristine and 1% LSO samples charged to 4.8 V. i) Li+ diffusion coefficient during the charge–discharge process for the
pristine and 1% LSO samples.

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (7 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

with increasing Li2SnO3 content, because the Li2SnO3 is electro- the spinel phase can be induced by the Li2SnO3 nanocoating,
chemically inactive whereas Li2MnO3 is partially activated due leading to the formation of 3D Li+ transport channels within
to the presence of the nanocoating layer during the first cycle. the spinel structure, which promote the rapid diffusion of Li+ at
However, the initial Coulombic efficiency (≈92%) of the 1% LSO high charge and discharge rates. Third, during the coating and
is higher than that of the pristine sample (≈72%). Moreover, synchronous lithiation process, trace amounts of Sn can diffuse
the specific capacity of the nanocoated samples are improved into the bulk phase from the nanocoating, thereby expanding
by the third cycle. In particular, the 1% LSO has a specific the interslab spacing and promoting Li+ migration.[24] Finally,
capacity of up to 315 mAh g−1. In comparison, the capacity of the excellent capacity is partly derived from the reversible
the pristine material has decreased to 225 mAh g−1. All of these anionic redox activity that is activated by the nanocoating-
observations indicate that the reversible redox properties of induced Sn doping and spinel structure. We further elucidate
the material have been improved in the nanocoated samples. this point with the detailed results of soft XAS and DSC.
Cyclic voltammograms (CVs) were recorded for the pristine To investigate the predominant rate capability in more
and 1% LSO samples (Figure S7, Supporting Information). As detail, the cyclic performances of the pristine and nanocoated
shown in Figure S7a of the Supporting Information, the two materials were measured at 0.2 C, 1 C, and 5 C between 2
oxidation peaks of the pristine material, occurring at ≈4.0 and and 4.8 V and at room temperature, as shown in Figure 5c–e.
≈4.6 V, can be ascribed to the oxidation of Ni2+ and the extrac- The 1% LSO sample retains a specific capacity of 220 mAh g−1
tion of Li+ from the layered structure with oxygen release, after 100 cycles, with a Coulombic efficiency of 99% at 0.2 C.
respectively. In the cathodic scan, the two distinct reduction In comparison, the specific capacity of the pristine material is
peaks at 3.75 and 3.3 V can be attributed to the reduction of only 177 mAh g−1 after 100 cycles, and the remaining capacity
Ni4+ and Mn4+. The current associated with the reduction peaks is only 90% of its initial value. The specific capacity of 3%
increases in the second and third cycles, which means that a LSO is better than that of the pristine sample, which retains
continuous activation process has been established in the ini- 196 mAh g−1 and is consistent with the rate property. The high
tial cycles. However, as the cycle number is increased, the peak rate capacities of 3% and 5% LSO are worse than those of 1%
at ≈4.6 V becomes weaker, indicating the irreversible release LSO and the pristine material because of the uneven nature of
of oxygen. The CV curves of 1% LSO (Figure S7b, Supporting the coating and impure SnO2 phase. The specific capacity of the
Information) show some distinct differences. Three obvious 1% LSO is 190 mAh g−1 with a discharge capacity of 98% after
peaks are present at ≈3 and ≈4.7 V during the oxidation pro- 200 cycles at the current density of 1 C. In contrast, the pristine
cess, corresponding to the oxidation of Mn3+ and Ni2+ from the material delivers a specific capacity of 138 mAh g−1, showing
spinel phase. As the scan is recorded for the reduction process, a capacity maintenance of 89%, as shown in Figure 5d. At the
three apparent peaks develop due to the high voltage spinel current density of 5 C, the 1% LSO exhibits the best cycle sta-
phase. The redox peaks of Ni cations at ≈4 V and the removal bility with a capacity retention of 98.2% after 500 cycles, as
of Li+ at 4.6 V are not obvious due to the overlap of the sloped shown in Figure 5e. Moreover, the cycle stability test conducted
region and the small amount of oxygen released.[10c] In addi- at the current density of 1 C at 55 °C is shown in Figure 5f. The
tion, it should be noted that in the first three cycles, the CV specific capacity of each sample type sharply increases during
curves of 1% LSO change only slightly, which further confirms the initial cycles, which may be caused by the more thorough
the structural stability and high reversibility of 1% LSO. activation and more vigorous kinetics of Li+ at this compara-
Figure 5b shows the rate performances of the four samples. tively high test temperature relative to room temperature.[25]
The charge–discharge current densities are recorded for 0.1, Interestingly, the discharge capacity of all materials is substan-
0.2, 0.5, 1, 2, and 5 C. The 1% LSO sample demonstrates the tially improved at 55 °C in comparison with that of room tem-
highest capacity at these different current densities, recorded at perature. For example, the capacity of the 1% LSO sample is
approximately 255, 233, 208, 188, 164, and 125 mAh g−1 from ≈300 mAh g−1 at the fifth cycle, with a remaining capacity of
0.1 to 5 C, respectively. Collectively, these results are better 235.1 mAh g−1 after 70 cycles at 1 C. This may be due to faster
than those recorded (212, 184, 166, 141, 123, and 99 mAh g−1) Li+ kinetics at 55 °C, more anions in the redox reaction, and the
for the pristine material. The 5% LSO performs the worst of enhanced layered structure stability associated with the Li2SnO3
the four samples, which is consistent with the characteristics nanocoating layer. Moreover, the electrochemical performance
of the initial charge–discharge curves. This poor behavior can of 1% LSO was observed for a series of Li1.2Mn0.6Ni0.2O2, as
be attributed to the presence of an impurity phase (SnO2) and shown in Table S3 (Supporting Information). The excellent
the uneven nature of the coating, as shown in Figure 1 and cycle performance of this material can be ascribed to the syner-
Figure S3 (Supporting Information). It is noted that the specific gistic effect of Li2SnO3 nanocoating-induced Sn doping and the
capacity of 1% LSO remains the highest and is the most stable formation of the spinel phase heterostructure during synchro-
when the current density returns to 0.1 C after 5 C, due to nous lithiation; this synergy enhances the anion redox activity
its high charge–discharge rate, which is associated with the and stabilizes the surface lattice oxygens.
stability of its structure. This superior rate capability can be According to the above analysis, we find that the 1% LSO
ascribed to the following points. First, the uniform nanocoating sample shows the best performance in terms of the rate and
can be formed by a small amount of Li+ conductor derived from cycle stability of all four samples. There is almost no reduction
the Li2SnO3 layer during the synchronous lithiation process, in the discharge capacity of 1% LSO at the current densities of
which is beneficial to the rapid migration of the Li ion; simi- 0.2, 1, and 5 C. It is worth noting that voltage fading is a critical
larly, this nanocoating can protect the layered structure from problem in Li-rich Mn-based oxide cathode materials. Figure 5g
electrolyte corrosion and the evolution of oxygen gas. Second, shows the average voltage decrease of the four samples over

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (8 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

200 cycles at the current density of 1 C. Clearly, the discharge Accordingly, we can clearly see that the DLi value of the 1%
average voltage plateau of the 1% LSO is higher than that of LSO is superior to that of the other materials, whether during
the others, with less fading than the pristine material. This phe- charging or discharging. This result is in agreement with
nomenon can be observed at the different discharge rates, as the analysis presented above. The EIS values of all samples
shown in Figure S8 (Supporting Information). This result can are shown in Figure 6. Figure 6a,b present Nyquist plots of
be attributed to the lower degree of cation mixing and disso- the samples before cycling, and Figure 6c,d shows the plots
lution, the mitigation of oxygen release and the enhanced sta- after 200 cycles. From this comparison, we can determine that
bility of the structure relative to that of the pristine material, the resistance of the 1% LSO sample is the smallest of the four
due to the nanocoating-induced doping and the development of samples, which indicates the side reaction that typically occurs
the spinel phase. between the electrode surface and electrolyte in 1% LSO has
The excellent cycle stability is closely related to the mate- been prevented by the nanocoating. As a result, the nanocoated
rial’s structure. To confirm the presence of lattice oxygen and layer has stabilized the lattice oxygen and improved the electro-
the thermal stability of the material, DSC was carried out.[6b] chemical performance of the material.
Figure 5h shows the DSC profiles of the pristine and 1% LSO In addition, the in situ and ex situ XRD patterns of the pris-
samples upon charging to 4.8 V. The exothermic peak of the tine and 1% LSO samples under different charge and discharge
pristine sample appears at 287 °C with a high amount of heat conditions are shown in Figure 7. The (003) peaks first shift to
generation (371.7 J g−1). In contrast, the main exothermic peak the left and then to the right during charging and discharging;
of 1% LSO appears at 293 °C with a lower degree of heat release in this analysis, the movement of the peak corresponding to the
(207.2 J g−1). These results indicate that the 1% LSO sample has pristine material is larger than that of 1% LSO. Moreover, the
a better thermal stability and more stable lattice oxygen, which changes in unit cell volume of the pristine and 1% LSO mate-
may be ascribed to the protective ability of the nanocoating and rials were derived from refined ex situ XRD data and are plotted
the reduced TMO covalence caused by Sn doping. As a result, in Figure 7g. It can be seen that the variation in cell volume
both the cycle performance and the safety property have been of the 1% LSO sample is much smaller than that of the pris-
significantly improved. tine sample, which is primarily attributed to the nanocoating-
To further understand the excellent electrochemical perfor- induced Sn doping and development of the spinel phase.[11a]
mance of 1% LSO, potentiostatic intermittent titration tech- The (311) and (101) peaks of the 1% LSO sample were observed
nique (PITT) and electrochemical impedance spectroscopy by in situ XRD, as shown in Figure 7h, where both peaks
(EIS) were conducted, as shown in Figures 5i, S9 (Supporting exhibit only small shifts in position. Collectively, these results
Information), and Figure 6. The Li+ diffusion coefficient can be indicate that the structure is stable because of the small vari-
determined based on the following formula, used with PITT ation in its unit cell volume and the associated property of a
analysis[26] reversible redox reaction.
To further reveal the origin of the improved specific capacity
DLi = − d ln (I ) /dt ][
⋅ 4L2 /π 2  and stability of 1% LSO relative to the pristine material, O

Figure 6.  EIS Spectra of the four cathode materials. a) EIS spectra of the four samples before cycling. b) Magnified image of (a) before cycling. c) EIS
spectra of the four samples after 200 cycles. d) Magnified image of (c) after 200 cycles.

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (9 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 7.  XRD Patterns and changes in cell volume for the pristine material and 1% LSO sample. a) Ex situ XRD spectra of the pristine material.
b, c) Magnified images of (a) at 17.5–19.6° and 36–37.25°, respectively. d) Ex situ XRD spectra of the 1% LSO sample. e, f) Magnified images of (d) at
17.5–19.6° and 36–37.25°, respectively. g) Cell volume evolution of the pristine and 1% LSO samples at different charge–discharge states. h) Selected
in situ XRD peaks, (311) and (101), for 1% LSO during the first cycle.

K-edge ex situ XAS of both materials was conducted. As is during oxygen redox in the Li2SnO3-coated sample relative to
shown in Figure  8a,b, the 3d band pre-edge peak at ≈529 eV that of the pristine sample; this result is consistent with the
shifts to a lower energy during the charging process, which is excellent cyclic performance observed for 1% LSO.[4g,h] More-
attributed to the oxidation of TM ions. Of interest is the fact over, the eg peak (≈531 eV) of the pristine sample is more criti-
that the pre-edge peak did not return to its original position cally distorted than that of the 1% LSO sample at 4.8 V. This
when discharging to 2 V in the pristine material. In contrast, distortion may be related to the structural rearrangement
the pre-edge peak of the 3d band of the 1% LSO sample shifted during oxygen evolution, which further indicates that the lattice
to a higher energy after discharging to 2 V, which indicates oxygen of the 1% LSO sample is more stable than that of the
the reversible nature of the charge compensation mechanism pristine sample.[27]

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (10 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

of the Supporting Information shows the XPS spectra of the


pristine and 1% LSO samples, respectively. The red area rep-
resents the existence of O2− 2 species at 530.5 eV.[4e,28] The
larger areas of the 1% LSO sample demonstrate the higher
oxygen redox activity. Furthermore, to exclude any interfer-
ence from the surface coating layer, both the pristine and 1%
LSO electrode samples charged at 4.8 V were also etched at
50 and 100 nm, respectively, and then were analyzed by XPS. As
shown in Figure S10b,d of the Supporting Information, the red
areas of the etched samples of 1% LSO (50 and 100 nm, respec-
tively) are also larger than those of the pristine samples etched
at the same thickness, which indicates the higher oxygen redox
activity of 1% LSO than the pristine sample. The above analysis
confirms that both the oxygen redox reactivity and its revers-
ibility are enhanced in the 1% LSO sample.

3. Conclusion
The oxygen redox reactivity and reversibility of LMLO were
enhanced by applying an Li2SnO3 nanocoating layer, which
induced the formation of a spinel phase and Sn doping during
the synchronous lithiation process. The crystal structure
of the material was studied by XRD and neutron diffraction.
Our analyses revealed that the 1% LSO sample exhibited a
lower degree of cation mixing, a more stable structure, and
a lower degree of volumetric expansion compared with the
pristine material and other Li2SnO3-coated samples. HRTEM
analysis revealed the spatial distribution of the Li2SnO3 nano-
coating and spinel phase. The nanocoating promoted the sta-
bility of surface lattice oxygens, as well as Li+ diffusion, and
Figure 8. Ex situ soft XAS at different voltage states in FY Mode.
STEM revealed the doping of Sn atoms within the TM layers
a,b) represent O K-edge soft XAS spectra collected at different states of and an expanded interslab spacing due to the pillar structure.
charge for the pristine and 1% LSO samples, respectively. c) A comparison Soft XAS spectral analysis clarified that the 1% LSO sample
of the relative integrated intensity for the pristine and 1% LSO samples in possessed a weaker covalence of TMO and an enhanced
the lower energy region (shaded regions in (a) and (b)). reversible oxygen redox reaction, both of which resulted in
a better electrochemical performance than that seen from
The integrated intensity of low energy peaks with different others. The specific discharge capacities for 1% LSO were 222,
charge states can supply important information about the unoc- 190, and 122 mAh g−1 at the current densities of 0.2, 1, and
cupied hole states of the oxygen 2p orbitals, was previously 5 C under room temperature, respectively, and the Coulombic
verified with soft XAS by Bruce and and co-workers and Yoon efficiency was ≈99%. The maximum specific capacity was as
et al.[4g,h,19b] The integrated intensity of the pre-edge peak was high as ≈300 mAh g−1, even at 55 °C. This synergistic strategy
calculated between 525 and 534 eV, and the results are shown combines the advantages of a nanocoating, Sn doping, and
in Figure 8c (the shaded region). By comparing the variation spinel phase formation and may open the door to significant
of the integrated intensity of the O K-edge when using the FY gains in our understanding and the design and synthesis of
mode applied to the pristine and 1% LSO samples at different high-performance cathode materials with enhanced reversible
charge states, the peak intensity of 1% LSO is seen to be greater anionic redox properties.
than that of the pristine material, as shown in Figure 8c; this
observation indicates that more oxygen redox chemistry occurs
in the 1% LSO sample.[4g,h,19b,27] It is worth noting that the
integrated intensity of the 1% LSO sample returns closer to 4. Experimental Section
the starting state than that of the pristine material when dis- Material Synthesis: The precursor was synthesized by our previous
charging to 2 V, which further indicates the better reversibility method.[11a] To synthesize Mn0.6 Ni0.2 CO3 @SnO2, the precursor (1 g)
of the charge compensation behavior of the oxygen redox was first dissolved in absolute ethanol (10 mL). According to the weight
activity in the 1% LSO sample relative to that of the pristine of the precursor, stoichiometric amounts of SnCl4 dissolved in solution
were dropped into the corresponding precursor solution. Then, the
material. To further confirm the oxygen redox activity and
mixed solution was dried overnight at 80 °C after vigorously stirring
degree of reversibility, we also performed ex situ XPS of the at 700–800 rpm for 12 h to mix thoroughly. A stoichiometric amount
electrodes under different charge–discharge states (as shown of LiOH·H2O (5% excess) was mixed into the precursor, and then the
in Figure S10 of the Supporting Information). Figure S10a,c mixture was calcined at 900 °C for 12 h (5 °C min−1). The obtained

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (11 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

samples were labeled as pristine, 1% LSO, 3% LSO, and 5% LSO Keywords


according to the respective amount of Sn in each.
Material Characterization: The crystal structures of the as-prepared Li-rich cathode, lithium-ion batteries, neutron diffraction, oxygen redox,
samples were characterized using XRD (Ultima IV, Cu Kα) in the 2θ soft X-ray absorption spectra
range of 10–80°. Data were recorded at a step size of 0.02° and a scan
rate of 10° per minute. Neutron diffraction data were collected at the Received: September 21, 2018
research reactor BER II at Helmholtz-Zentrum Berlin für Materialien Revised: December 5, 2018
und Energie (HZB) using the FIREPOD powder diffractometer (E9). Published online: January 7, 2019
Scanning electron microscopy (SEM, Hitachi SU-8010) and TEM
(Tecnai G2 F30) were used to observe the morphologies of the samples
and to obtain detailed lattice spacing measurements of the phases
within the samples. Aberration-corrected scanning transmission [1] a) J. Lee, D. A. Kitchaev, D. H. Kwon, C. W. Lee, J. K. Papp,
electron microscopy JEM ARM200F (JEOL, Tokyo, Japan) was Y. S. Liu, Z. Lun, R. J. Clement, T. Shi, B. D. McCloskey,
performed, aided by two CEOS (CEOS, Heidelberg, Germany) probe J. Guo, M. Balasubramanian, G. Ceder, Nature 2018, 556, 185;
aberration correctors, to probe the structure at the atomic scale. XPS b) M. S. Whittingham, Chem. Rev. 2004, 104, 4271; c) C. Zhan,
was measured using a Thermo Escalab 250Xi with a monochromatic T. Wu, J. Lu, K. Amine, Energy Environ. Sci. 2018, 11, 243;
X-ray source (E (Al Kα)  = 1486.6 eV). Each spectrum was normalized d) J. B. Goodenough, K. S. Park, J. Am. Chem. Soc. 2013, 135, 1167;
using the region of the C 1s (284.6 eV) peak as the line position of e) R. Schmuch, R. Wagner, G. Hörpel, T. Placke, M. Winter, Nat.
adventitious carbon. Soft XAS of powdered material was performed
Energy 2018, 3, 267; f) M. Li, J. Lu, Z. Chen, K. Amine, Adv. Mater.
at the Russian–German Beamline of the synchrotron Bessy II, Berlin,
2018, 30, 1800561.
Germany. Data were obtained in both total electron yield (TEY) and
fluorescence yield (FY) modes at room temperature under ultrahigh [2] J. Kim, H. Cho, H. Y. Jeong, H. Ma, J. Lee, J. Hwang, M. Park,
vacuum (10−9 Torr). The total number of electrons emitted from the J. Cho, Adv. Energy Mater. 2017, 7, 1602559.
sample were counted in TEY mode, and the energy-dispersive Bruker [3] a) B. Li, D. Xia, Adv. Mater. 2017, 29, 1701054; b) J. L. Shi,
XFlash fluorescence detector was used in FY mode. Soft XAS was D. D. Xiao, M. Ge, X. Yu, Y. Chu, X. Huang, X. D. Zhang, Y. X. Yin,
performed as an ex situ experiment at beamline BL08U1-A at the X. Q. Yang, Y. G. Guo, L. Gu, L. J. Wan, Adv. Mater. 2018, 30,
Shanghai Synchrotron Radiation Facility (SSRF). 1705575.
Electrochemical Measurements: The positive electrodes were formed [4] a) G. Assat, J. M. Tarascon, Nat. Energy 2018, 3, 373; b) S. K. Jung,
from the active materials (80 wt%), carbon black (10 wt%), and K. Kang, Nat. Energy 2017, 2, 912; c) C. Zhan, Z. Yao, J. Lu,
polyvinylidene fluoride (10 wt%), mixed in N-methyl pyrrolidinone. L. Ma, V. A. Maroni, L. Li, E. Lee, E. E. Alp, T. Wu, J. Wen, Nat.
The slurry was deposited onto an Al foil and then dried under vacuum Energy 2017, 2, 963; d) A. Grimaud, W. T. Hong, Y. Shao-Horn,
at 120 °C for 12 h. The mass of the active materials loaded on each J. M. Tarascon, Nat. Mater. 2016, 15, 121; e) M. Sathiya, G. Rousse,
electrode was ≈2 mg. All electrode cells were assembled in a glovebox K. Ramesha, C. P. Laisa, H. Vezin, M. T. Sougrati, M. L. Doublet,
under an Ar atmosphere with oxygen and water contents of less than D. Foix, D. Gonbeau, W. Walker, A. S. Prakash, M. B. Hassine,
0.1 ppm. The PITT and cyclic voltammograms were performed on L. Dupont, J. M. Tarascon, Nat. Mater. 2013, 12, 827; f) E. McCalla,
an electrochemical workstation (Metrohm-Autolab, PGSTAT 302N) A. M. Abakumov, M. Saubanère, D. Foix, E. J. Berg, G. Rousse,
instrument. EIS was performed on an electrochemical workstation M. L. Doublet, D. Gonbeau, P. Novák, G. Van Tendeloo,
(Metrohm-Autolab, PGSTAT 302N) instrument with an amplitude of
Science 2015, 350, 1516; g) K. Luo, M. R. Roberts, N. Guerrini,
5 mV and a frequency range from 100 kHz to 0.1 Hz. Galvanostatic
N. Tapia-Ruiz, R. Hao, F. Massel, D. M. Pickup, S. Ramos, Y. S. Liu,
charge–discharge cycling was performed between 2.0 and 4.8 V (vs Li/
J. Guo, A. V. Chadwick, L. C. Duda, P. G. Bruce, J. Am. Chem. Soc.
Li+) by using a battery tester (NEWARE) at different current densities.
2016, 138, 11211; h) K. Luo, M. R. Roberts, R. Hao, N. Guerrini,
D. M. Pickup, Y. S. Liu, K. Edstrom, J. Guo, A. V. Chadwick,
L. C. Duda, P. G. Bruce, Nat. Chem. 2016, 8, 684.
Supporting Information [5] E. Hu, X. Yu, R. Lin, X. Bi, J. Lu, S. Bak, K. W. Nam, H. L. Xin,
C. Jaye, D. A. Fischer, K. Amine, X. Q. Yang, Nat. Energy 2018, 3,
Supporting Information is available from the Wiley Online Library or 690.
from the author. [6] a) M. Saubanère, E. McCalla, J. M. Tarascon, M. L. Doublet, Energy
Environ. Sci. 2016, 9, 984; b) X. D. Zhang, J. L. Shi, J. Y. Liang,
Y. X. Yin, J. N. Zhang, X. Q. Yu, Y. G. Guo, Adv. Mater. 2018, 30,
1801751.
Acknowledgements [7] a) A. J. Perez, Q. Jacquet, D. Batuk, A. Iadecola, M. Saubanère,
This work was supported by the National Natural Science Foundation of G. Rousse, D. Larcher, H. Vezin, M. L. Doublet, J. M. Tarascon,
China (Grant No. 11575192), the Natural Science Foundation of Beijing Nat. Energy 2017, 2, 954; b) P. E. Pearce, A. J. Perez, G. Rousse,
(Grant No. 2182082), the Scientific Instrument Developing Project M. Saubanere, D. Batuk, D. Foix, E. McCalla, A. M. Abakumov,
(Grant No. ZDKYYQ20170001), the International Partnership Program G. Van Tendeloo, M. L. Doublet, J. M. Tarascon, Nat. Mater. 2017,
(Grant No. 211211KYSB20170060), the Strategic Priority Research 16, 580.
Program (Grant No. XDB28000000), and “The Hundred Talents Project” [8] P. K. Nayak, E. Levi, J. Grinblat, M. Levi, B. Markovsky,
of the Chinese Academy of Sciences. The authors also thank the staff N. Munichandraiah, Y. K. Sun, D. Aurbach, ChemSusChem 2016, 9,
at the BL08U1-A beamline of SSRF for their support. The allocation of 2404.
beamtime at RGBL beamline, BESSY-II, Berlin, Germany, is gratefully [9] A. Manthiram, K. Chemelewski, E. S. Lee, Energy Environ. Sci. 2014,
acknowledged. 7, 1339.
[10] a) F. Jiao, J. Bao, A. H. Hill, P. G. Bruce, Angew. Chem., Int. Ed.
2008, 47, 9711; b) M. Lin, L. Ben, Y. Sun, H. Wang, Z. Yang, L. Gu,
X. Yu, X. Q. Yang, H. Zhao, R. Yu, Chem. Mater. 2015, 27, 292;
Conflict of Interest c) N. H. Vu, J. C. Im, S. Unithrattil, W. B. Im, J. Mater. Chem. A
The authors declare no conflict of interest. 2018, 6, 2200.

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (12 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[11] a) J. Zhang, R. Gao, L. Sun, Z. Li, H. Zhang, Z. Hu, X. Liu, Phys. Z. Li, Z. Hu, X. Liu, Phys. Chem. Chem. Phys. 2016, 18,
Chem. Chem. Phys. 2016, 18, 25711; b) D. Luo, G. Li, C. Fu, 13322.
J. Zheng, J. Fan, Q. Li, L. Li, Adv. Energy Mater. 2014, 4, 1400062; [17] P. Oh, S. M. Oh, W. Li, S. Myeong, J. Cho, A. Manthiram, Adv. Sci.
c) Q. Xia, X. Zhao, M. Xu, Z. Ding, J. Liu, L. Chen, D. G. Ivey, 2016, 3, 1600184.
W. Wei, J. Mater. Chem. A 2015, 3, 3995; d) D. Wang, I. Belharouak, [18] Z. Ding, M. Xu, J. Liu, Q. Huang, L. Chen, P. Wang, D. G. Ivey,
G. Zhou, K. Amine, Adv. Funct. Mater. 2013, 23, 1070; e) F. Wu, W. Wei, ACS Appl. Mater. Interfaces 2017, 9, 20519.
N. Li, Y. Su, H. Shou, L. Bao, W. Yang, L. Zhang, R. An, S. Chen, [19] a) B. Li, H. Yan, J. Ma, P. Yu, D. Xia, W. Huang, W. Chu, Z. Wu, Adv.
Adv. Mater. 2013, 25, 3722; f) Y. Pei, Q. Chen, Y. C. Xiao, Funct. Mater. 2014, 24, 5112; b) W. S. Yoon, M. Balasubramanian,
L. Liu, C. Y. Xu, L. Zhen, G. Henkelman, G. Cao, Nano Energy K. Y. Chung, X. Q. Yang, J. McBreen, C. P. Grey, D. A. Fischer,
2017, 40, 566; g) R. Yu, X. Zhang, T. Liu, L. Yang, L. Liu, Y. Wang, J. Am. Chem. Soc. 2005, 127, 17479; c) J. Xu, M. Sun, R. Qiao,
X. Wang, H. Shu, X. Yang, ACS Appl. Mater. Interfaces 2017, 9, S. E. Renfrew, L. Ma, T. Wu, S. Hwang, D. Nordlund, D. Su,
41210. K. Amine, J. Lu, B. D. McCloskey, W. Yang, W. Tong, Nat. Commun.
[12] a) U.-H. Kim, S.-T. Myung, C. S. Yoon, Y.-K. Sun, ACS Energy Lett. 2018, 9, 947.
2017, 2, 1848; b) Y. Liu, D. Ning, L. Zheng, Q. Zhang, L. Gu, [20] Y. Ye, J. E. Thorne, C. H. Wu, Y. S. Liu, C. Du, J. W. Jang, E. Liu,
R. Gao, J. Zhang, A. Franz, G. Schumacher, X. Liu, J. Power Sources D. Wang, J. Guo, J. Phys. Chem. B 2018, 122, 927.
2018, 375, 1; c) X. Li, K. Zhang, D. Mitlin, Z. Yang, M. Wang, [21] S. Shen, Y. Hong, F. Zhu, Z. Cao, Y. Li, F. Ke, J. Fan, L. Zhou, L. Wu,
Y. Tang, F. Jiang, Y. Du, J. Zheng, Chem. Mater. 2018, 30, 2566; P. Dai, M. Cai, L. Huang, Z. Zhou, J. Li, Q. Wu, S. Sun, ACS Appl.
d) Q. Liu, X. Su, D. Lei, Y. Qin, J. Wen, F. Guo, Y. A. Wu, Y. Rong, Mater. Interfaces 2018, 10, 12666.
R. Kou, X. Xiao, Nat. Energy 2018, 3, 936; e) Y. Wang, Z. Yang, [22] a) C. Yin, H. Zhou, Z. Yang, J. Li, ACS Appl. Mater. Interfaces 2018,
Y. Qian, L. Gu, H. Zhou, Adv. Mater. 2015, 27, 3915; f) X. Zeng, 10, 13625; b) W. Xiao, C. Xin, S. Li, J. Jie, Y. Gu, J. Zheng, F. Pan,
C. Zhan, J. Lu, K. Amine, Chem 2018, 4, 690; g) Y. Zhao, J. Li, J. Mater. Chem. A 2018, 6, 9893.
J. R. Dahn, Chem. Mater. 2017, 29, 5239; h) J. Huang, H. Liu, T. Hu, [23] Z. Wei, Y. Gao, L. Wang, C. Zhang, X. Bian, Q. Fu, C. Wang, Y. Wei,
Y. S. Meng, J. Luo, J. Power Sources 2018, 375, 21; i) S. Hildebrand, F. Du, G. Chen, Chem. - Eur. J. 2016, 22, 11610.
C. Vollmer, M. Winter, F. M. Schappacher, J. Electrochem. Soc. [24] M. Sathiya, A. M. Abakumov, D. Foix, G. Rousse, K. Ramesha,
2017, 164, A2190; j) J. Kim, M. Noh, J. Cho, H. Kim, K. B. Kim, M. Saubanere, M. L. Doublet, H. Vezin, C. P. Laisa, A. S. Prakash,
J. Electrochem. Soc. 2005, 152, A1142; k) S. Lee, Y. Cho, H. K. Song, D. Gonbeau, G. VanTendeloo, J. M. Tarascon, Nat. Mater. 2015, 14,
K. T. Lee, J. Cho, Angew. Chem., Int. Ed. 2012, 51, 8748; l) J. Wang, 230.
X. Sun, Energy Environ. Sci. 2012, 5, 5163; m) G. Assat, A. Iadecola, [25] B. Xiao, H. Liu, J. Liu, Q. Sun, B. Wang, K. Kaliyappan, Y. Zhao,
D. Foix, R. Dedryvère, J. M. Tarascon, ACS Energy Lett. 2018, 3, M. N. Banis, Y. Liu, R. Li, T. K. Sham, G. A. Botton, M. Cai, X. Sun,
2721. Adv. Mater. 2017, 29, 1703764.
[13] S. Kim, M. Aykol, V. I. Hegde, Z. Lu, S. Kirklin, J. R. Croy, [26] Z. Y. Li, R. Gao, L. Sun, Z. Hu, X. Liu, J. Mater. Chem. A 2015, 3,
M. M. Thackeray, C. Wolverton, Energy Environ. Sci. 2017, 10, 2201. 16272.
[14] Y. L. Ding, J. Xie, G. S. Cao, T. J. Zhu, H. M. Yu, X. B. Zhao, Adv. [27] S. Hy, W. N. Su, J. M. Chen, B. J. Hwang, J. Phys. Chem. C 2012, 116,
Funct. Mater. 2011, 21, 348. 25242.
[15] X. Zhou, L. Yu, X. W. Lou, Adv. Energy Mater. 2016, 6, 1600451. [28] a) X. Li, Y. Qiao, S. Guo, Z. Xu, H. Zhu, X. Zhang, Y. Yuan, P. He,
[16] a) Q. Q. Qiao, L. Qin, G.-R. Li, Y.-L. Wang, X.-P. Gao, J. Mater. M. Ishida, H. Zhou, Adv. Mater. 2018, 30, 1705197; b) Q. Li, Y. Qiao,
Chem. A 2015, 3, 17627; b) J. Zhang, H. Zhang, R. Gao, S. Guo, K. Jiang, Q. Li, J. Wu, H. Zhou, Joule 2018, 2, 1134.

Adv. Funct. Mater. 2019, 29, 1806706 1806706  (13 of 13) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like