You are on page 1of 8

Communication

www.advmat.de

Electron-State Confinement of Polysulfides for Highly


Stable Sodium–Sulfur Batteries
Chao Ye, Yan Jiao, Dongliang Chao, Tao Ling, Jieqiong Shan, Binwei Zhang, Qinfen Gu,
Kenneth Davey, Haihui Wang,* and Shi-Zhang Qiao*

with metal–sulfur batteries is the “shuttle


Confinement of polysulfides in sulfur cathodes is pivotal for eliminating the effect.” This arises from solubility of the
“shuttle effect” in metal–sulfur batteries, which represent promising solu- intermediate metal polysulfides, together
tions for large-scale and sustainable energy storage. However, mechanistic with a parasitic reaction between the poly-
exploration and in-depth understanding for the confinement of polysulfides sulfides and the metal anodes. This leads
to a low Coulombic efficiency (CE) and a
remain limited. Consequently, it is a critical challenge to achieve highly
rapid capacity decay in metal–sulfur bat-
stable metal–sulfur batteries. Here, based on a 2D metal–organic framework teries.[3] For example, room-temperature
(2D MOF), a new mechanism to realize effective confinement of polysulfides sodium–sulfur (RT Na–S) batteries benefit
is proposed. A combination of in situ synchrotron X-ray diffraction, elec- from low cost of Na compared with the
trochemical measurements, and theoretical computations reveal that the other metal–sulfur batteries. However, its
dynamic electron states of the Ni centers in the 2D MOF enable the interac- practical application is severely hindered
by the shuttle effect because of the high
tion between polysulfides and the MOF in the discharge/charge process to
reactivity of Na.[4] Therefore confinement
be tuned, resulting in both strong adsorption and fast conversion kinetics of polysulfides in sulfur cathodes is critical
of polysulfides. The resultant room-temperature sodium–sulfur batteries are to long-term stability and practical applica-
amongst the most stable reported so far, thus demonstrating that the new tion of metal–sulfur batteries.[2a,5]
mechanism opens a promising avenue for the development of high-perfor- Currently, polysulfides confinement
with weak physical adsorption on carbon-
mance metal–sulfur batteries.
based materials would inevitably inhibit
their conversion kinetics, resulting in
a compromise between high discharge
As a promising electrode material, sulfur benefits from low cost capacity and low capacity decay of the sulfur cathodes.[6] On
and high theoretical specific capacity of ≈1675 mAh g−1.[1] Addi- the other hand, too strong polysulfides adsorption would cause
tionally, sulfur cathodes can be conjugated with a range of metal decomposition of the polysulfides, leading to blockage of the
anodes in metal–sulfur batteries with high energy density of adsorption sites.[7] The keys for efficient polysulfides confine-
≈350 W h kg−1 (based on devices). This holds promise for prac- ment lie in appropriate adsorption of the polysulfides and their
tical energy-storage applications.[2] However, a major drawback facilitated conversion kinetics, which are largely determined
by local electronic state of a sulfur cathode material.[8] For
example, on N-doped graphene, pyridinic N with an extra pair
C. Ye, Dr. Y. Jiao, Dr. D. Chao, J. Shan, Dr. B. Zhang, Dr. K. Davey, of electrons interacts with terminal Li in lithium polysulfides
Prof. S.-Z. Qiao to form Li bond, which greatly promotes lithium polysulfides
School of Chemical Engineering and Advanced Materials confinement.[9] However, systematic investigation to correlate
The University of Adelaide
Adelaide, SA 5005, Australia local electron state with adsorption behaviors of a series of
E-mail: s.qiao@adelaide.edu.au polysulfides remains very limited, which is due to the com-
Prof. T. Ling, Prof. S.-Z. Qiao plexity of the polysulfide conversion process. For example, in
School of Materials Science and Engineering the case of the RT Na–S batteries, at least four sodium poly-
Tianjin University sulfides (NaPoSs) including Na2S5, Na2S4, Na2S2, and Na2S are
Tianjin 300072, China
involved.[10] Consequently, realizing efficient polysulfides con-
Dr. Q. Gu
Australian Synchrotron (ANSTO)
finement in sulfur cathodes has been a challenging topic in
800 Blackburn Rd, Clayton, VIC 3168, Australia high-energy-density metal–sulfur batteries.
Prof. H. Wang Although polysulfides are too sensitive to be detected
School of Chemistry and Chemical Engineering in air, advances in in situ synchrotron-based characteriza-
South China University of Technology tions permit the identification of specific polysulfides and the
Guangzhou 510640, China tracking of dynamic macroscopic polysulfides conversions.
E-mail: hhwang@scut.edu.cn
In-depth comprehension can thus be offered for macroscopic
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adma.201907557.
polysulfides conversion kinetics.[11] Nevertheless, atomic-level
understanding for the polysulfides adsorption behaviors is
DOI: 10.1002/adma.201907557 imperative and remains difficult to achieve experimentally.[12]

Adv. Mater. 2020, 1907557 1907557  (1 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Extraordinary progress in density functional theory (DFT) A hydrothermal method was adopted to fabricate a single
computations that take into account the surface chemistry and crystal bulk Ni2(PymS)4 MOFs (Ni-MOF-bulk, PymSH = 2-mer-
local electron state of the sulfur cathode materials is essential captopyrimidine).[15] Morphology of the Ni-MOF-bulk was
for investigating the polysulfides adsorption behaviors.[13] For characterized by scanning electron microscopy (SEM) to reveal
instance, d-orbital electron numbers of the transitional metal its layered structure (Figure S1, Supporting Information).
sulfides strongly determine adsorption energies of lithium High-angle annular dark-field scanning transmission-elec-
polysulfides.[14] Therefore, combination of the advanced in situ tron microscopy (HAADF-STEM) images and corresponding
synchrotron characterizations and computational quantum energy-dispersive spectroscopy (EDX) elemental maps revealed
chemistry can reveal the mechanism of polysulfides confine- a uniform distribution of C, N, S, and Ni (Figure S2, Sup-
ment on cathode materials. One can engineer potential sulfur porting Information). Mechanical exfoliation was used on the
cathode materials with efficient polysulfides confinement by Ni-MOF-bulk via wet ball-milling in acetone. Because of weak
tailoring their electron states. It is thus urgently required for interaction between its layers, the Ni-MOF-bulk was readily
introducing the aforementioned advanced methodology to the exfoliated into 2D MOF nanosheets (Ni-MOF-2D).[16] This was
field of metal–sulfur batteries. evidenced by a broad signal from 3100 to 3150 cm−1 in the
Here, we report a new mechanism for polysulfides con- Raman spectrum that is assigned to N···H hydrogen bonds
finement based on a 2D metal–organic framework (2D between layers of the Ni-MOF-bulk (Figure S3, Supporting
MOF). Using the MOF as a model together with a judicial Information). By contrast, a negligible corresponding signal in
combination of in situ synchrotron X-ray diffraction (XRD), the Ni-MOF-2D would imply breaking of the N···H hydrogen
electrochemical tests, and DFT computations, we demonstrate bonds as a result of exfoliation.[17] The schematic shows this
that dynamic electron states of Ni centers in the MOF enable layered structure of Ni-MOF-2D with a (002) interlayer-space
tuning of interaction between polysulfides and the MOF in of 0.84 nm (Figure  1a). XRD patterns demonstrate that the
the discharge/charge process, facilitating strong polysulfide Ni-MOF-2D maintains the original crystalline phase of the Ni-
adsorption and rapid polysulfides conversion kinetics. The MOF-bulk (Figure 1b). The inset high-resolution transmission
resulting performance of the RT Na–S batteries is amongst the electron microscopy (HRTEM) image shows an outstanding
best ones reported so far. We report for the first time the cor- (002) peak in Ni-MOF-2D pattern to evidence successful exfo-
relation between the local electron states of the sulfur cathode liation of the bulk MOF.[15] Representative SEM and trans-
materials and facilitated polysulfides conversion kinetics. Our mission electron microscopy (TEM) measurements were per-
findings offer a practical means to address the shuttle effect in formed to aid visualization of the morphology of Ni-MOF-2D.
metal–sulfur batteries and guide the design of sulfur cathode These showed a 2D nanosheet structure (inset of Figure 1c;
materials for highly stable metal–sulfur batteries. Figures S4a and S5a, Supporting Information). Atomic force

Figure 1.  Structural characterization and electron-state analyses of Ni-MOF-2D nanosheets. a) Schematic of Ni-MOF-2D with few-layer structure from
side-view and top-view, in which the blue, light-gray, yellow, purple, and white spheres represent, respectively, N, C, S, Ni, and H. b) XRD patterns of
Ni-MOF-bulk and Ni-MOF-2D. Inset is HRTEM image of the few-layer Ni-MOF-2D. c) HRTEM and d) HAADF-STEM image and corresponding EDX
mapping images of Ni-MOF-2D. The inset of (c) is a representative TEM image of Ni-MOF-2D. e) Ni L-edge, f) S K-edge, g) C K-edge, and h) N K-edge
NEXAFS spectra of Ni-MOF-2D and Ni-MOF-bulk. Insets show the enlarged regions for charge transfer information.

Adv. Mater. 2020, 1907557 1907557  (2 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

microscopy images confirmed this 2D morphology with a CE held at around 100%. As is seen in Figure 2b, S/Ni-MOF-2D
thickness of ≈8.6 nm (Figure S4b, Supporting Information). exhibited a series of advantageous discharge capacities of 516,
HRTEM images (Figure 1c) together with the selected-area- 416, 372, 331, and 284 mAh g−1 when cycled at, respectively,
electron-diffraction pattern (Figure S5b, Supporting Informa- 0.1, 0.2, 0.5, 1, and 2 C (1 C = 1675 mA g−1). When current
tion) gave measured lattice distances of 0.23 and 0.27 nm, density was switched back to 0.2 C a high discharge capacity of
corresponding, respectively, to the (244) and (241) facets of 406 mAh g−1 was maintained. By contrast, S/Ni-MOF-bulk and
Ni2(PymS)4 MOFs. To provide a more detailed composition of S/conductive carbon cathodes showed poor rate performance
Ni-MOF-2D, HAADF-STEM image with corresponding EDX with limited capacities under high rate (Figure S12, Supporting
elemental maps is presented as Figure 1d. This shows that the Information). As is shown in Figures S13 and S14 following
C, N, S, and Ni elements are uniformly distributed and con- further cycling at 0.2 C, the S/Ni-MOF-2D exhibited an ultralow
firms that the Ni-MOF-2D maintains the physical phase of pris- capacity decay of 0.024% per cycle (406 to 313 mAh g−1 in
tine Ni-MOF-bulk. 970 cycles). This was significantly better than that for either
Because it is reported that an optimized electron-state of S/Ni-MOF-bulk (0.074%) or S/conductive carbon (0.095%).
cathode materials is favorable for improving discharge capacity The intact electrode structure of the S/Ni-MOF-2D compared
and cycling stability of sulfur cathodes.[18] We performed DFT with the other two electrodes after cycling further confirms the
calculations and conducted Bader charge analysis to predict the best cycling stability among the three electrodes (Figure S15,
charge redistribution from the Ni-MOF-bulk to the Ni-MOF- Supporting Information). This excellent cycling stability of
2D.[19] Findings show that ≈10% of electrons from S transfer to S/Ni-MOF-2D is superior to those of the metal oxides, metal
Ni in processing of Ni-MOF-bulk to Ni-MOF-2D (Figure S6 and sulfides, or hybrid sulfur cathodes reported recently (Figure 2c;
Table S1, Supporting Information). Synchrotron-based near- Table S2, Supporting Information). In addition, we prepared a
edge X-ray absorption fine structure (NEXAFS) characteriza- sulfur electrode with higher sulfur loading of ≈2.6 mg, which
tions were performed on both Ni-MOF-bulk and Ni-MOF-2D to exhibits capacity of 244 mAh g−1 and CE of around 100% after
investigate any structural impact on electron transfer behavior. 200 cycles under 0.2 C, indicating the high promise of this
NEXAFS is a powerful tool to provide reliable information on strategy to promote the future practical application of Na–S bat-
local electron-states. According to the white-line adsorption teries (Figure S16, Supporting Information).
energies, Figure 1e and inset present a clear view of the Ni Cyclic voltammetry (CV) tests under a low scan rate of
L-edge spectra in which Ni-MOF-2D exhibits increased electron 0.1 mV s−1 between 0.5 and 2.8 V were performed to investigate
density following exfoliation of Ni-MOF-bulk. For the S K-edge, electrochemical behavior of sulfur and electrochemical evolu-
the peak at around 2473.0 eV can be assigned to S–Ni coordi- tion of NaPoSs from S/Ni-MOF-2D cathode (Figure 3a). During
nation species.[20] The increased intensity of the signal in Ni- the initial cathodic scan, two prominent peaks centered at
MOF-2D compared with that in Ni-MOF-bulk demonstrate a around 2.08 and 0.73 V were observed. These corresponded to
decreased electron density of S following exfoliation (Figure 1f). the solid–liquid transition from elemental sulfur to Na2Sx (long-
X-ray photoelectron spectra measurements were performed chain NaPoSs, 6 ≤ x ≤ 8) and a further reduction to less-soluble
on these two elements of the Ni-MOF-bulk and Ni-MOF-2D to Na2S2 and Na2S. In the ensuing cathodic scan, two strong
substantiate electron transfer behavior (Figure S7, Supporting cathodic peaks at 1.59 and 0.94 V were observed that were
Information).[21] Compared with Ni-MOF-bulk, Ni-MOF-2D highly reproducible. These are related to reduction of Na2Sx.
showed a lower Ni valence state and a higher S valence state, For the anodic scan, only one reproducible peak at 2.04 V was
which is consistent with the computational results and indicates observed. This corresponds to oxidation of Na2S2 and Na2S into
the electron transfer from S to Ni. For the C and N K-edge, the Na2Sx. The lower potential gap between first reduction and oxi-
peak at around 287.0 and 399.2 eV can be assigned to CNC dation peaks in S/Ni-MOF-2D suggests lower polarization and
species (Figure 1g,h). The increased intensity of this feature in faster conversion kinetics of the sodium polysulfides compared
C K-edge and decreased intensity of the feature in N K-edge with those for S/Ni-MOF-bulk (Figure S17, Supporting Infor-
implies an electron transfer from C to N.[22] Overall, these find- mation).[24] Additionally, we estimated the voltage gaps (ΔE) at
ings confirm a charge redistribution during the exfoliation 50% depth of discharge. These are closely related to electrode
from Ni-MOF-bulk to Ni-MOF-2D. This can be attributed to the polarization (Table S3, Supporting Information). S/Ni-MOF-2D
breaking of weak interlayer interactions (e.g., N***H bonds), exhibited a lower ΔE2D than either S/Ni-MOF-bulk or S/con-
which is suggested by the decrease of corresponding Raman ductive carbon. This finding indicates significant enhancement
signal after exfoliation (Figure S3, Supporting Information).[23] of NaPoSs redox kinetics on the Ni-MOF-2D and is consistent
To investigate the electrochemical performance of the as-pre- with the CV observations.[25]
pared Ni-MOFs in RT Na–S batteries we assembled three sulfur To further investigate the NaPoSs’ electrochemical and
electrodes of S/Ni-MOF-2D, S/Ni-MOF-bulk and S/conductive adsorption behavior on Ni-MOF-2D, in situ synchrotron XRD
carbon using active materials with a sulfur mass ratio of 48.6% measurements were conducted in transmission mode. A modi-
(Figure S8, Supporting Information). Long-term cycling experi- fied 2032-type coin cell of S/Ni-MOF-2D was performed using
ments at a high rate of 1 C and 2 C were conducted to investi- an in-house design (Figure S18, Supporting Information).[26]
gate electrochemical performance of S/Ni-MOF-2D (Figure 2a; As can be seen in Figure 3b during the initial discharge from
Figures S9–S11, Supporting Information). High capacities of ≈2.2 V, two peaks at 20.2° and 12.9° were observed in the XRD
347 and 241 mAh g−1 were maintained after 1000 continuous patterns. These are assigned to organized layers of soluble
cycles under 1 C and 2 C that refer respectively to low capacity Na2Sx with relative ordered structure.[27] When the battery was
decay of 0.042% and 0.052% per cycle, together with a stabilized discharged to ≈1.6 V, three new peaks emerged at 11.9°, 20.1°,

Adv. Mater. 2020, 1907557 1907557  (3 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

a 1000

100
800

Colulombic efficiency (%)


S/Ni-MOF-2D
Capacity (mAh g-1) S/conductive carbon 80
600 S/Ni-MOF-bulk

60
400
40

200
20

0 0
0 100 200 300 400 500 600 700 800 900 1000
Cycle
b 1000 S/Ni-MOF-2D
c 0.02
S/conductive carbon
S/Ni-MOF-bulk This work
800

Capacity decay (cycle-1, %)


0.04
[S5]
Capacity (mAh g )

0.06
-1

0.1C

[S10]
600 0.08 [S11]
0.1
0.2C

[S12]
0.2C

0.5C

[S13]
400
1C

0.2
2C

[S14]
[S8]
[S15] [S9]
200 0.4 [S7]
[S16]
0.6 [S6]
0.8
0 1
5 10 15 20 25 30 0 300 600 900 1200
Cycle Cycle number

Figure 2. Electrochemical performance of S/Ni-MOF-2D in RT Na–S batteries. a) Cycling performances and Coulombic efficiencies (CE) of the
S/Ni-MOF-2D, S/Ni-MOF-bulk, and S/conductive carbon at 1 C after 0.1 C activation. b) Rating capacities of the three sulfur electrodes. c) Comparison
of cycle numbers and capacity retentions of recently reported RT Na–S batteries with the current work in which deeper color refers to greater current
density. References for [S5] etc. are given in the Supporting Information.

a b 12
charge

0.1 2.04 V
10

0.0 8
Current (mA)

Time (h)

1.59 V 2.08 V 6
-0.1
0.94 V
4

-0.2
2
0.73 V
discharge
0
0.5 1.0 1.5 2.0 2.5 3.0 3 2 1 0 -110 11 12 13 20 21
Voltage (V, vs Na +/Na) Voltage (V) 2 Theta

Figure 3.  Electrochemical behavior of NaPoSs on Ni-MOF-2D. a) CV curves of S/Ni-MOF-2D at scan rate of 0.1 mV S−1. b) Initial galvanostatic dis-
charge/charge curve and corresponding in situ synchrotron XRD patterns of S/Ni-MOF-2D in which the colored patterns indicate major changes along
the discharge/charge process; the yellow and cyan spheres represent S and Na atoms, respectively.

Adv. Mater. 2020, 1907557 1907557  (4 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

and 20.6°, corresponding, respectively, to (131), (341), and (302) density for Ni-MOF-2D-Na2S5 is demonstrated by the increased
facets of Na2S5 (JCPDF no. 77–0294). This finding reveals that white-line adsorption energy of Ni L-edge (Figure 4b). In the
the peak in the CV scan at 1.59 V is a result of the reduction of C K-edge, no electron gain or loss is observed (Figure 4c). It
Na2Sx to Na2S5. The peak at 11.6° is derived from the (220) facet is therefore concluded that this interaction leads to decreased
of Na2S4 (JCPDF no. 71–0516) and was generated when the bat- electron density on S and N and is responsible for the conju-
tery was discharged further to ≈1.0 V. This indicates that the gation effect in the heterocyclic linker rings. This is different
peak at 0.94 V in the CV curves can be reliably assigned to to reported open-metal sites in 3D MOFs in which electron
the formation of Na2S4. This peak for Na2S4 faded into a new density would be expected to increase due to interaction with
(200) peak for Na2S2 (JCPDF no. 81–1771) at 12.1° when the S in the lithium polysulfides.[33] Our results imply therefore
battery was further discharged to 0.8 V. The final discharge to that electrons of Ni transferred to NaPoSs through the linker.[34]
≈0.7 V revealed a peak at 10.6°. This is attributed to the (111) More importantly, even lower Ni electron densities in Ni-MOF-
facet of Na2S (JCPDF no. 77–2149). In the charge from 0.5 to 2D-Na2S2 and Ni-MOF-2D-Na2S are demonstrated by higher
2.8 V, most of the Na2S in the battery was oxidized to Na2Sx. It white-line adsorption energy of Ni L-edge compared with that in
is notable that broader peaks are related to Na2S2, Na2S4, and Ni-MOF-2D-Na2S5 (Figure S22, Supporting Information). This
Na2S5 when compared with narrow peaks for Na2Sx. This is indicates that the electron state of the Ni centers is dynamic
attributed to the formation of an organized Na2S4 and Na2S5 during discharge/charge processes. By contrast, Ni-MOF-bulk
layer that is the result of strong adsorption on Ni-MOF-2D.[28] exhibited weak interaction with Na2S5 as is evidenced in the
Importantly, overall results from the combination of the in situ Raman and NEXAFS spectra (Figure S23, Supporting Informa-
synchrotron XRD and CV measurements underscore the facili- tion). To confirm NaPoSs adsorption capabilities of Ni-MOF-
tated conversion kinetics of all NaPoSs on Ni-MOF-2D. 2D, Ni-MOF-bulk, and the linker PymSH, samples were treated
To study the origin of the electrochemical performance of with a 5 × 10−3 m Na2S5 solution for 1 h. As is illustrated in
Ni-MOF-2D, we used PymSH and another Ni-based MOFs Figure S24 of the Supporting Information, the Ni-MOF-2D pos-
(Ni-BDC MOFs) as controls and investigated the individual sesses the highest adsorption capabilities based on same weight
roles of the heterocyclic linker and the Ni center (Figures S19 or same surface area in comparison with Ni-MOF-bulk, con-
and S20, Supporting Information). Results showed that both ductive carbon, and PymSH. Overall, these results underscore
PymSH (without the Ni) and Ni-BDC MOFs (without the that NaPoSs are adsorbed on Ni-MOF-2D through interaction
PymSH ligand) are inactive in sulfur cathodes. To provide an with the heterocyclic linker as an electron donor and NaPoSs
objective view of electrochemical performance of Ni-MOF-2D, as an electron acceptor. The N/S sites of Ni-MOF-2D provide
we evaluated capacities normalized by electrochemically active effective polysulfides adsorption sites which are lacking on the
surface areas (ECSAs) on S/Ni-MOF-2D, S/Ni-MOF-bulk, and conductive carbon. This is the reason that the S/Ni-MOF-2D
S/conductive carbon cathodes (Figure S21, Supporting Infor- significantly outperforms S/conductive carbon. Notably, elec-
mation).[29] S/Ni-MOF-2D exhibted the greatest ECSA together tron transfer behaviors are also observed on Ni centers, which
with greatest normalized capacity under 0.2 C, 1 C, and 2 C might contribute to the interaction between the polysulfides
(Table S4, Supporting Information). The former finding is and Ni-MOF-2D.
attributed to the more exposed active sites of the 2D mor- To further gain atomic-level insights into the polysulfides
phology. The latter demonstrates a highly efficient NaPoSs confinement on Ni-MOF-2D, we performed theoretical calcula-
confinement on Ni-MOF-2D from strong chemical adsorption tions. The (002) facet was determined as an exposed surface of
and fast redox kinetics toward NaPoSs.[30] It is concluded there- the Ni-MOF-2D based on the HRTEM images. Ab initio mole-
fore that the linkers and Ni centers synergistically contribute cular dynamics (AIMD) calculations were applied to investigate
to NaPoSs adsorption and conversion kinetics on Ni-MOF-2D. NaPoSs adsorption energies on Ni-MOF-2D. The energy pro-
To explore the mechanism of NaPoSs adsorption on Ni-MOF- file (Ea) of all five Na2Sn molecules (n = 1, 2, …, 5) adsorbed on
2D, a series of spectra surveys were carried out. Raman spectra the Ni-MOF-2D surface (Ni-MOF-2D-Na2Sn) were evaluated for
of Ni-MOF-2D and Na2S5-treated Ni-MOF-2D (Ni-MOF-2D- 4000 fs (Figure  5a). AIMD simulations predicted Ea increases
Na2S5) were investigated to study the change in the chemical with the length of the Na2Sn chain as ≈−4.4, −4.0, −3.9, −2.6,
environment of the heterocyclic linker. As is shown Figure 4a, and −2.5 eV, respectively. These high adsorption energies indi-
right inset, stretching bands (ν ring) of heterocyclic linker ring cate strong adsorption of the NaPoS on Ni-MOF-2D when com-
were observed in the range of 1450–1650 cm−1. These are good pared with those of reported materials in sulfur cathodes for
indicators of the electron density of the linker.[31] The bands cen- lithium/sodium polysulfides adsorption.[6] DFT calculations
tered at 1565 and 1541 cm−1 in the Ni-MOF-2D actually down- were carried out to optimize the Ni-MOF-2D-Na2Sn configura-
shift to 1563 and 1533 cm−1 following treatment with Na2S5. tions for analyses of electron transfer induced by the adsorption
This finding suggests that the heterocyclic linker especially (Figure S25, Supporting Information). The results implies that
electron-rich N interacts with Na2S5 via electron transfer.[32] N and S can function as a dominant adsorption site to interact
Importantly, the left inset shows that the CS bond stretching with Na in Na2Sn and the N/S can interact with Na in Na2Sn
vibration band (ν CS) at 1167 cm−1 in Ni-MOF-2D upshifts as electron donors and acceptors through Na–N/S interaction.
to 1179 cm−1 in Ni-MOF-2D-Na2S5. This finding indicates a This finding agrees well with both the Raman and NEXAFS
decrease in electron density on S following interaction with results. Importantly, as is shown in Figure 5b, with Na2Sn
Na2S5. NEXAFS characterizations were carried out on Ni-MOF- adsorption energies increasing from Ni-MOF-2D-Na2S5 to Ni-
2D-Na2S5 to investigate the change in local electron density. As MOF-2D-Na2S, electron states of the Ni centers tend to change
is shown in the regions of Ni L-edge, a decreased Ni electron from an electron accummulation state (in Ni-MOF-2D-Na2S5)

Adv. Mater. 2020, 1907557 1907557  (5 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

a
Ni-MOF-2D C-S ring
Ni-MOF-2D-Na2S5
1167 1563
1565
1179
Intersity (a.u.) 1541
1533

1140 1170 1200 1230 1500 1530 1560 1590

600 800 1000 1200 1400 1600 1800


Raman shift (cm-1)
b c Ni-MOF-2D
Ni L-edge Ni-MOF-2D C K-edge
Ni-MOF-2D-Na2S5 Ni-MOF-2D-Na2S5

Normalizaed adsorption (a.u.)


Normalizaed adsorption (a.u.)

852.5 853.0 853.5

845 850 855 860 865 870 875 280 290 300 310 320
Photon energy (eV) Photon energy (eV)

Figure 4.  Analysis of electron transfer between Ni-MOF-2D and Na2S5. a) Raman spectra, b) Ni L-edge NEXAFS spectra, and c) C K-edge NEXAFS
spectra of Ni-MOF-2D and Ni-MOF-2D-Na2S5.

to an electron depletion state (in Ni-MOF-2D-Na2S). This capacities and capacity retention. By contrast, the Ni-MOF-bulk
indicates that electron transfer from Ni to N correlates with suffers from weak NaPoSs confinement (Figure S26, Sup-
the strengthened Na–N/S interaction and increased adsorp- porting Information). Density of states analysis revealed a sem-
tion energy from Na2S5 to Na2S on Ni-MOF-2D, which veri- iconducting property of Ni-MOF-2D with a smaller band gap
fies the NEXAFS results of Ni L-edge (Figure S22, Supporting of ≈0.56 eV compared with those of reported MOFs in sulfur
Information). cathodes (Figure S27, Supporting Information). The semicon-
The schematic illustrates effect of the Ni centers’ electron ducting property of Ni-MOF-2D is believed responsible for the
states on NaPoSs confinement (Figure 5c). Basically, the Ni enhanced electron transfer between Ni centers and the linkers.
centers with high redox capability are understood to strengthen All these results demonstrate that the Ni centers of S/Ni-MOF-
the Na–N/S interaction through electron transfer from the Ni- 2D exhibited dynamic electron states during charge/discharge
MOF-2D to the NaPoSs. This strengthened Na–N/S interac- process, which evokes tuned Na (polysulfide)-N/S (Ni MOF)
tion should facilitate the sodiation process from Na2S5 to Na2S interaction. Therefore, strong polysulfides adsorption and fast
during discharge. During charge, the weakened Na–N/S inter- polysulfides conversion kinetics can be realized and efficient
actions also facilitate the desodiation process of Na2S to Na2S5 polysulfides confinement was promoted on the S/Ni-MOF-2D.
due to electron transfer from the NaPoSs to the Ni-MOF-2D. We conclude that our work demonstrates a new mechanism
Therefore, this dynamic electron transfer between Ni-MOF-2D for effective polysulfides confinement on a 2D Ni (II) MOF
and NaPoSs facilitates NaPoSs sodiation–desodiation kinetics and its chemical origin. The electrochemical performances of
on Ni-MOF-2D. The Ni centers with high redox capability the resulting sulfur cathodes are shown to be superior to those
appear to result from the charge redistribution of the Ni-MOF- of all sulfur cathode materials in RT Na–S batteries. On the
2D.[35] As a result, enhanced NaPoSs adsorption and conver- basis of synchrotron-based in situ XRD, NEXAFS characteriza-
sion kinetics on Ni-MOF-2D significantly boosts the discharge tions, electrochemical tests, and systematic DFT computations,

Adv. Mater. 2020, 1907557 1907557  (6 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

a 4 b Ni-MOF-2D-Na2Sn
-5 conductive carbon-Na2Sn

Adsorption energy (eV)


-4
Adsorption energy (eV)

N site
-3

-4
-2
Ni center
S site
-1
-8
Ni-MOF-2D-Na2S
conductive carbon-Na2S
0
0 1000 2000 3000 4000 Na2S5 Na2S4 Na2S3 Na2S2 Na2S
MD time (fs)

c Discharge process Charge process


Stronger Na-N/S interaction
Strengthened Na-N/S interaction Weakened Na-N/S interaction
Weaker Na-N/S interaction

Solvated Na+

NaPoSs

Electron paths

Ni-MOF-2D

e- e-
e-

e-

Figure 5.  Computational investigation of NaPoSs confinement on Ni-MOF-2D. a) Adsorption energies of Na2Sn on graphene and Ni-MOF-2D as a
function of ab initio molecular dynamics (AIMD) simulation time (fs). b) Adsorption energies of Na2Sn with insets of charge difference analyses, in
which yellow and cyan isosurface represent electron accumulation and electron depletion. The color code is the same as for Figure 1a and the cyan
spheres represent Na atoms. The isosurface value is 0.005 e Å−3. c) Schematic of NaPoSs confinement on Ni-MOF-2D.

we demonstrated that strong polysulfides adsorption and fast Supporting Information


polysulfides conversion kinetics can be realized by tuning
Supporting Information is available from the Wiley Online Library or
Na–N/S interaction via dynamic electron states of Ni centers,
from the author.
thus results in efficient polysulfides confinement on the Ni
MOF. We conclude that our findings offer a practical means for
production of highly stable metal–sulfur batteries and for the
design of electrode materials for broader applications in energy Acknowledgements
storage and conversion.
This work was supported financially by the Australian Research Council
(ARC) through Discovery Project and Linkage Project Programs
(DP160104866, LP160100927, and FL170100154) and by the National
Experimental Section Natural Science Foundation of China (51722103). C.Y. was supported
by the Chinese CSC Scholarship Program. The authors thank Dr. Bruce
Experimental details can be found in the Supporting Information. Cowie at the Australian Synchrotron for help with NEXAFS.

Adv. Mater. 2020, 1907557 1907557  (7 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advmat.de

Conflict of Interest [14] X. Chen, H.-J. Peng, R. Zhang, T.-Z. Hou, J.-Q. Huang, B. Li,


Q. Zhang, ACS Energy Lett. 2017, 2, 795.
The authors declare no conflict of interest. [15] Y. Zhao, M. Hong, Y. Liang, R. Cao, W. Li, J. Weng, S. Lu, Chem.
Commun. 2001, 11, 1020.
[16] Y. Yao, Z. Lin, Z. Li, X. Song, K.-S. Moon, C.-p. Wong, J. Mater.
Chem. 2012, 22, 13494.
Keywords [17] R. Yadav, D. Swain, P. P. Kundu, H. S. Nair, C. Narayana,
confinement of polysulfides, metal–organic frameworks, sodium–sulfur S. Elizabeth, Phys. Chem. Chem. Phys. 2015, 17, 12207.
batteries, sulfur cathodes, 2D materials [18] D. Liu, C. Zhang, G. Zhou, W. Lv, G. Ling, L. Zhi, Q. H. Yang, Adv.
Sci. 2018, 5, 1700270.
Received: November 18, 2019 [19] W. Tang, E. Sanville, G. Henkelman, J. Phys.: Condens. Matter 2009,
Revised: January 3, 2020 21, 084204.
Published online: [20] A. V. Soldatov, A. N. Kravtsova, M. E. Fleet, S. L. Harmer, J. Phys.:
Condens. Matter 2004, 16, 7545.
[21] a) F. Pei, L. Lin, D. Ou, Z. Zheng, S. Mo, X. Fang, N. Zheng, Nat.
Commun. 2017, 8, 482; b) H.-L. Zhang, S. D. Evans, J. R. Henderson,
[1] Y. Yang, G. Zheng, Y. Cui, Chem. Soc. Rev. 2013, 42, 3018. R. E. Miles, T. Shen, J. Phys. Chem. B 2003, 107, 6087; c) Z. Jiang,
[2] a) M. Salama, Rosy, R. Attias, R. Yemini, Y. Gofer, D. Aurbach, H. Xie, S. Wang, X. Song, X. Yao, H. Wang, Adv. Energy Mater. 2018,
M. Noked, ACS Energy Lett. 2019, 4, 436; b) X. Hong, J. Mei, 8, 1801433; d) H. Xiang, J. Chen, Z. Li, H. Wang, J. Power Sources
L. Wen, Y. Tong, A. J. Vasileff, L. Wang, J. Liang, Z. Sun, S. X. Dou, 2011, 196, 8651.
Adv. Mater. 2019, 31, 1802822. [22] Y. Jiao, Y. Zheng, K. Davey, S.-Z. Qiao, Nat. Energy 2016, 1, 16130.
[3] A. Manthiram, Y. Fu, S.-H. Chung, C. Zu, Y.-S. Su, Chem. Rev. 2014, [23] S. Leikin, V. A. Parsegian, W. H. Yang, G. E. Walrafen, Proc. Natl.
114, 11751. Acad. Sci. USA 1997, 94, 11312.
[4] a) X. Xu, D. Zhou, X. Qin, K. Lin, F. Kang, B. Li, D. Shanmukaraj, [24] D. Chao, C. Zhu, M. Song, P. Liang, X. Zhang, N. H. Tiep, H. Zhao,
T. Rojo, M. Armand, G. Wang, Nat. Commun. 2018, 9, 3870; J. Wang, R. Wang, H. Zhang, H. J. Fan, Adv. Mater. 2018, 30,
b) B.-W. Zhang, T. Sheng, Y.-X. Wang, S. Chou, K. Davey, S.-X. Dou, 1803181.
S.-Z. Qiao, Angew. Chem., Int. Ed. 2019, 58, 1484. [25] J. Zhang, Z. Li, Y. Chen, S. Gao, X. W. Lou, Angew. Chem., Int. Ed.
[5] Z. W. Seh, Y. Sun, Q. Zhang, Y. Cui, Chem. Soc. Rev. 2016, 45, 5605. 2018, 57, 10944.
[6] Q. Pang, X. Liang, C. Y. Kwok, L. F. Nazar, Nat. Energy 2016, 1, [26] Q. Gu, J. A. Kimpton, H. E. Brand, Z. Wang, S. Chou, Adv. Energy
16132. Mater. 2017, 7, 1602831.
[7] X. Chen, T. Hou, K. A. Persson, Q. Zhang, Mater. Today 2019, 22, [27] Y.-X. Wang, J. Yang, W. Lai, S.-L. Chou, Q.-F. Gu, H. K. Liu, D. Zhao,
142. S. X. Dou, J. Am. Chem. Soc. 2016, 138, 16576.
[8] a) Z. Li, B. Y. Guan, J. Zhang, X. W. Lou, Joule 2017, 1, 576; [28] J. Conder, R. Bouchet, S. Trabesinger, C. Marino, L. Gubler,
b) J. Zhou, R. Li, X. Fan, Y. Chen, R. Han, W. Li, J. Zheng, B. Wang, C. Villevieille, Nat. Energy 2017, 2, 17069.
X. Li, Energy Environ. Sci. 2014, 7, 2715; c) B.-W. Zhang, T. Sheng, [29] a) B. You, N. Jiang, M. Sheng, M. W. Bhushan, Y. Sun, ACS Catal.
Y.-D. Liu, Y.-X. Wang, L. Zhang, W.-H. Lai, L. Wang, J. Yang, 2015, 6, 714; b) H. Li, Y. Tao, C. Zhang, D. Liu, J. Luo, W. Fan,
Q.-F. Gu, S.-L. Chou, Nat. Commun. 2018, 9, 4082. Y. Xu, Y. Li, C. You, Z.-Z. Pan, M. Ye, Z. Chen, Z. Dong, D.-W. Wang,
[9] T.-Z. Hou, W.-T. Xu, X. Chen, H.-J. Peng, J.-Q. Huang, Q. Zhang, F. Kang, J. Lu, Q.-H. Yang, Adv. Energy Mater. 2018, 8, 1703438.
Angew. Chem., Int. Ed. 2017, 56, 8178. [30] H. Jin, C. Guo, X. Liu, J. Liu, A. Vasileff, Y. Jiao, Y. Zheng, S.-Z. Qiao,
[10] a) X. Yu, A. Manthiram, ChemElectroChem 2014, 1, 1275; b) I. Kim, Chem. Rev. 2018, 118, 6337.
J.-Y. Park, C. Kim, J.-W. Park, J.-P. Ahn, J.-H. Ahn, K.-W. Kim, [31] Y. S. Pang, H. J. Hwang, M. S. Kim, J. Mol. Struct. 1998, 441, 63.
H.-J. Ahn, J. Electrochem. Soc. 2016, 163, A611. [32] G. N. R. Tripathi, M. Clements, J. Phys. Chem. B 2003, 107, 11125.
[11] a) Y. Yan, C. Cheng, L. Zhang, Y. Li, J. Lu, Adv. Energy Mater. 2019, 9, [33] J. Zheng, J. Tian, D. Wu, M. Gu, W. Xu, C. Wang, F. Gao,
1900148; b) X. Chen, W. He, L.-X. Ding, S. Wang, H. Wang, Energy M. H. Engelhard, J.-G. Zhang, J. Liu, Nano Lett. 2014, 14, 2345.
Environ. Sci. 2019, 12, 938. [34] T. Ling, P. Da, X. Zheng, B. Ge, Z. Hu, M. Wu, X.-W. Du, W.-B. Hu,
[12] Q. Zhang, Y. Wang, Z. W. Seh, Z. Fu, R. Zhang, Y. Cui, Nano Lett. M. Jaroniec, S.-Z. Qiao, Sci. Adv. 2018, 4, eaau6261.
2015, 15, 3780. [35] X. Liang, C. Y. Kwok, F. Lodi-Marzano, Q. Pang, M. Cuisinier,
[13] T. Z. Hou, X. Chen, H. J. Peng, J. Q. Huang, B. Q. Li, Q. Zhang, H. Huang, C. J. Hart, D. Houtarde, K. Kaup, H. Sommer, Adv.
B. Li, Small 2016, 12, 3283. Energy Mater. 2015, 6, 1501636.

Adv. Mater. 2020, 1907557 1907557  (8 of 8) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like