You are on page 1of 12

Chemical Engineering Journal 421 (2021) 129707

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Laser-induced micro-explosion to construct hierarchical structure as


efficient polysulfide mediators for high-performance
lithium-sulfur batteries
Chao Zhou a, *, Ming Li b, Min Hong a, Nantao Hu a, *, Zhi Yang a, Liying Zhang a, Yafei Zhang a, *
a
Key Laboratory of Thin Film and Microfabrication Technology (Ministry of Education), School of Electronics, Information and Electrical Engineering, Shanghai Jiao
Tong University, Dong Chuan Road No.800, Shanghai 200240, China
b
State Key Laboratory of Transducer Technology, Shanghai Institute of Microsystem and Information Technology, Chinese Academy of Sciences, Shanghai 200050,
China

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

Keywords: Rational design and scalable construction of sulfur cathodes with high areal capacity and suppressed shuttle
Laser-induced effect are challenging for real Li-S application. Herein, laser-induced micro-explosive techniques were firstly
ZnO nanoparticles demonstrated as high-efficient polysulfide mediator for Li-S batteries, in which unique hierarchical structure
Shuttle effect
(nitrogen-doped porous carbon (LNPCs) skeletons uniformly decorated with metal oxide nanoparticles) were
Catalytic conversion
Li-S batteries
designed and constructed. The as-assembled nitrogen-doped porous carbons decorated with ZnO nanoparticles
(LNPCs-ZnO) hybrids cathode perform exceptionally high discharge capacity of 1184.9 mAh g− 1 at 0.5C and
excellent rate discharge capacity of 716.1 mA h g− 1 at 3C, as well as a remarkable cycling performance (a low-
capacity decay rate of 0.31% per cycle at 0.5C upon 1000 cycles). Moreover, impressive areal sulfur loading up to
6.7 mg cm− 2 leads to a high initial areal capacity of 6.36 mAh cm− 2 and excellent cycling stability at 0.2C. The
DFT calculations propose that the intrinsic interaction and the rapid catalytical effects play the key role in
suppressing the shuttling effect of LiPSs. This work provides new insights for novel structure design and opens up
a potential route to construct Li-S batteries with high areal sulfur loadings for their practical applications.

1. Introduction the above-mentioned issues, multifarious strategies have been devel­


oped to restrain polysulfide shuttling [8].
Lithium-sulfur batteries (Li-S batteries) have been deemed as one of Carbonaceous materials, such as meso- /microporous carbon, [9,10]
the most prospective devices for the next-generation energy storage carbon nanotubes, [11–13] carbon fibers, [14–16] carbon sphere,
systems owing to their ultrahigh theoretical specific capacity (1675 mA [17,18] grapheme [10b], [19–21] and reduced graphene oxide [22–25],
h g− 1), [1] remarkable energy density (2600 W h kg− 1), [2] low cost [3] have been widely used as the sulfur host materials. These carbon-based
and less contamination. [4] However, there are many issues which still materials can not only provide large pore volume and substantially
plagued the wide practical applications, 1) intrinsic insulating nature of boost the electronic conductivity of the sulfur-based cathode but also
element sulfur [5] and final discharge products of Li2S2/Li2S result in can further alleviate the diffusion of polysulfide in the electrode system.
sluggish reaction kinetics;[6] 2) soluble LiPS intermediates formed [26–31] However, the weak trapping ability and fast capacity decay can
during discharge process and diffusion between cathode and anode, result from the feeble Van der Waals bond energy between nonpolar
those intermediates dissolved in the ether electrolyte and formed carbonaceous materials and polar polysulfides. [32] Although hetero­
obvious concentration fluctuation, resulting in “shuttle effect”;[7] 3) atom (such as N, S, F, P, and B elements) doping into the carbon struc­
insulated Li2S layer is easily formed on the surface of the Li metal anode ture can produce polar sites and enhance the adsorbing interaction with
during discharge process, leading to the S element and Li metal anode polysulfide, it does not substantially improve the polysulfide trapping
irreversible loss, which further causes the irreversible capacity attenu­ ability over long cycling due to the weak chemisorption. [33–39]
ation, large polarization and poor cycling stability. [8] In order to solve Alternatively, polar inorganic nanomaterials, including metal oxides,

* Corresponding authors.
E-mail addresses: chaozhou0721@sjtu.edu.cn (C. Zhou), hunantao@sjtu.edu.cn (N. Hu), yfzhang@sjtu.edu.cn (Y. Zhang).

https://doi.org/10.1016/j.cej.2021.129707
Received 15 January 2021; Received in revised form 20 March 2021; Accepted 3 April 2021
Available online 15 April 2021
1385-8947/© 2021 Elsevier B.V. All rights reserved.
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

[40–42] metal sulfides, [26,43–48] metal phosphides, [49–51] metal high-performance wearable and portable devices.
nitrides, [52–54] and metal carbides [55–57], etc., have widely been
demonstrated that adequate functional polar groups had strong inter­ 2. Experimental section
action with LiPSs and blocked the LiPSs diffusion. In this respect, con­
structing 3D porous carbonaceous hosts with polar inorganic 2.1. Preparation of 3D porous LNPCs-ZnO hybrid frameworks
nanomaterials is preferable in addressing the long cycling life and pol­
ysulfide shuttle issues. In light of recent research work, it was demon­ The LNPCs-ZnO hybrid frameworks were prepared by using the same
strated that ZnO NPs had an excellent LiPSs anchoring ability and laser scribing method proposed in our previous works. [68,86] Typi­
superior binding energy with LiPSs. [58–60] For example, Kumar et al. cally, 40 mmol Zn(NO3)2⋅6H2O and 40 mmol glycine were dissolved in
[59] proposed a novel interconnected conductive frameworks with 10 mL Dimethylacetamide (DMAc) solution with continuous magnetic
strongly coupled zinc oxide (ZnO) nanowires and macroporous nickel stirring at 65 ◦ C to form a uniform solution. And then 10 mL poly (amic
foam to obtain an areal sulfur loading of 3 mg cm− 2 and a reversible acid) (PAA) solution was also added into the as-obtained solution. PAA
discharge capacity of 776 mA h g− 1 at 1C after 200 cycles. Shi et al. [60] was prepared by the same method that had been described in our pre­
developed free-standing electrode materials combined with reduced vious work. The mixture solution was magnetically stirred for 2 h to
graphene oxide (rGO)-sulfur hybrid aerogel and atomic layer deposition form a homogeneous solution. [86] The PAA solutions containing Zn
(ALD) ZnO coatings for Li-S battery, and a discharge capacity of 998 mA (NO3)2⋅6H2O and glycine were then cast onto the surface of slides. The
h g− 1 at 0.2C was achieved. Despite the wide varieties of polar nano­ thickness of the gel coating contained PAA, glycine, and Zn(NO3)2 was
materials have been investigated, the rate performance and polysulfide controlled at 100 μm through the scraper. Subsequently, the coating was
trapping ability are still not satisfactory with a high areal sulfur loading cured on slides upon evaporation at 80 ℃ for 12 h, followed by laser
due to the poor electronic conductivity and low specific surface area of direct writing under ambient conditions using a CO2 laser cutter plat­
metal oxides. [29,61,62] Moreover, it is imperative to open up a novel form, which employed an infrared wave laser with a laser power of 25
path to overcome the above-mentioned problems. W, a wavelength of 10.6 μm (Trotec Speedy 100), and a beam size of
Since laser-induced graphene was firstly reported in 2014, [63] it has laser was ~ 100 μm. In this experiment, the laser power and scan rate
been aroused warm research interest in the field of energy storage giving were fixed at 4.8 W and 35 cm/s, respectively, during laser induction. In
credit to their high surface area, [64] superior electronic conductivity, order to remove the unreacted precursors and by-products of LNPCs-
[63,64] and excellent thermal stability. [63,65] These excellent physical ZnO hybrids, the LNPCs-ZnO hybrids were alternately washed with
and chemical properties make them proverbially studied in the field of ethanol and deionized water three times and then dried in the vacuum
Li-ion battery, [66] Na-ion battery, [67] etc. The Li/Na-ion battery oven at 60 ◦ C for 12 h. For comparison, LNPCs were prepared with a
based on pristine LIG or heteroatom-doped LIG exhibits excellent elec­ similar method without Zn moieties.
trochemical performances. To further enhance the discharge capacity
and rate performance of LIG-based Li/Na-ion battery, transition-metal 2.2. Preparation of LNPCs-ZnO/S cathode
oxides [68] and transition-metal sulfide [69] have been introduced to
form hybrids with LIG or heteroatom-doped LIG hybrid frameworks. In LNPCs-ZnO/S cathode was achieved by a feasible melt-diffusion
our previous work, [68] biomass gelatin solutions and manganese (II) method. [49] Briefly, LNPCs-ZnO hybrid frameworks were mixed and
phthalocyanine were used as precursors to construct 3D laser-induced ground with sublimed sulfur powder at a different weight ratio of 3:7,
MnO/Mn3O4/N-doped-graphene hybrid framework on copper foils 2:8, and 1:9, respectively. The obtained composites were transferred
through a simple one-step laser-inducing process, showing exceptionally into a 10 mL Teflon-lined autoclave and then followed by maintaining
reversible capacity of 992 mAh g− 1 at 0.2 A g− 1. Although great pro­ the temperature at 155 ◦ C for 12 h under argon atmospheres. For
gresses have been made in the field of Li/Na-ion batteries, most of these comparison, LNPCs/S was also obtained via a similar procedure, apart
researches are focusing on the electrical conductivity and reversible from LNPCs was used as the host material instead of LNPCs-ZnO.
capacity optimization. However, some problems still exist in terms of
cycle stability and rate performance. Until now, few works are evolving 2.3. Preparation of Li2S6 solution and adsorption test
in unique laser-induced heteroatom-doped porous carbon/ transition
metal oxides hybrids through a one-step laser-induced micro-explosive In this experiment, the Li2S6 solution was chosen as the representa­
technique for Li-S battery and it is difficult to rationally design a tive of polysulfides. To prepare the Li2S6 solution, Li2S and sulfur
unique polar-non-polar structure with excellent conductivity and high powders with a molar ratio of 1:5 were added into a certain amount of
specific surface area to strongly trap polysulfide as well as long-term 1,3-dioxolane (DOL) and 1,2-dimethoxyethane (DME) solution (1 : 1, v/
cycling stability. v) under vigorous stirring at 60 ◦ C for 24 h in the glove-box filled with
Herein, a facile laser-induced micro-explosive strategy was demon­ Ar. The concentration of the as-obtained Li2S6 solution was 5 mmol L-1.
strated to rationally design a novel 3D porous cathode configuration for Before the polysulfides adsorption test, all the LNPCs and LNPCs-ZnO
Li-S batteries based on N-doped porous carbon hierarchical frameworks, hybrids were dried at 60 ◦ C under vacuum for 12 h. The LNPCs and
which are simultaneously decorated with polysulfide-confining ZnO LNPCs-ZnO hybrids with equal weight (10 mg) were immersed in a 2 mL
NPs, hereafter referred to as LNPCs-ZnO. The gel coatings, containing Li2S6 solution undergoing static adsorption for 12 h, respectively. The
poly (amic acid) (PAA), glycine, and Zn(NO3)2, were cast onto the sur­ color variation of supernatant with time can be observed and the
face of glass slides, followed by undergoing graphitization during the adsorption capability of LNPCs and LNPCs-ZnO to Li2S6 was tested by
micro-explosive process via a facile laser irradiation at room tempera­ UV–vis spectroscopy. The blank Li2S6 solution was used as the reference.
ture. As a result, unique 3D porous architectural LNPCs-ZnO hybrid
frameworks were obtained. As a cathode in Li-S batteries, the unique 2.4. Material characterization
structure of LNPCs-ZnO hybrid frameworks not only exhibited high
reversible capacity, but also alleviated polysulfide shuttling effect and The morphologies of the above-mentioned materials were charac­
greatly improved cycling stability. Moreover, giving credit to the high terized by FE-SEM analysis (Carl Zeiss Ultra55). TEM, HRTEM and EDX
specific surface area and high electronic conductivity of these unique mapping were all performed on a JEOL 2010HT microscope operating at
structures, high sulfur utilization and high-rate performance of Li-S 200 keV. X-ray photoelectron spectroscopy (XPS) was performed on a
battery were achieved. This study will indeed provide a new direction Kratos Axis Ultra DLD and fitted with XPS peak 4.1 software. The
to construct unique 3D porous hybrid frameworks applied in Li-S bat­ structural information of the samples was conducted on XRD, utilizing
tery-based devices, which holds great potentials for next-generation Cu Kα as a radiation source and 40 kV as working voltage. Raman

2
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

spectra were measured by a Renishaw in Via Reflex Raman spectrometer layer of 15 Å was added in the Z direction to avoid slab interlayer
with a 532 nm laser. The UV–Vis absorption spectra were obtained from possible interaction as shown in Figure A larger surface area was
a Perkin-Elmer Lambda 950 UV–Vis–NIR spectrophotometer. N2 generated to provide favorable adsorption locations for Li2Sx species
adsorption–desorption isotherms were obtained at 78 K using a Quan­ (These Li2Sx species include S8, Li2S, Li2S2, Li2S4, Li2S6, and Li2S8).
tachrome Autosorb-3B surface analyzer relying on the Brunauer- An all-electron method with the generalized gradient approximation
Emmett-Teller (BET) method. The pore size distributions were ac­ (GGA) was exploited to execute the spin-polarized DFT computations.
quired according to the Barrett-Joyner-Halenda (BJH) method. The The model of Perdew-Burke-Ernzerhof (PBE) functional in the DMol3
sulfur content in these hybrids was performed by thermogravimetric Code was implemented.1-4 Also, the double numerical plus polarization
Analysis (TGA) from 30 to 500 ◦ C with a programmed heating rate of (DNP) basis set was used. A convergence criterion of 10-5 a.u. on the
5 ◦ C/min under nitrogen atmospheres. total energy, maximum force of 0.002 a.u-1, and maximum displacement
of 0.005 Å were used for self-consistent field (SCF) calculations. The
2.5. Li2S nucleation measurements binding energy (E) was defined by the following equation:

Ebinding = E(ZnO (0 0 2)/Li2Sx))-E(ZnO -E(Li2Sx) (2)


The mixture of S and Li2S at a molar ratio of 7:1 was added in 1 M (0 0 2)

lithium bis(trifluoromethylsulfonyl)imide (LiTFSI) in a solvent of DOL/ where the E(ZnO (0 0 2)/Li2Sx)), E(ZnO (0 0 2), E(Li2Sx) are, respectively, the
DME (1:1, v/v) with stirring at 60 ◦ C for 24 h in glovebox. 20 µL Li2S8 total energies of the selected (0 0 2) surface, the ZnO NPs, and the Li2Sx
(0.5 mol/L) catholyte was added onto LNPCs-ZnO, while 20 µL 1.0 M (1 ≤ x ≤ 8) molecule adsorbed on the surface.
LiTFSI (DME/DOL, v/v = 1:1) electrolyte onto Li anode. The assembly
cell was galvanostatically discharged to 2.06 V at a current of 0.112 mA 3. Results and discussion
and then kept potentiostatically at 2.05 V for Li2S nucleation and growth
until the current dropped below 10-5 A. The construction workflow of LNPCs-ZnO hybrid frameworks for Li-S
batteries are schematically demonstrated in Fig. 1. For the purpose of
2.6. Electrochemical measurements directly obtaining the LNPCs-ZnO hybrid frameworks, the precursor
solution containing PAA, glycine, and Zn(NO3)2 was cast onto the sur­
The detailed electrochemical performance of the LNPCs/S, and face of slides (Fig. 1(a)). The thickness of the gel coating was fixed at
LNPCs/S cathodes was measured using a CR2025 coin cell, assembled in 100 μm through the scraper. A solid coating attached on slides was able
an Ar-filled glove box (H2O<0.1 ppm and O2<0.1 ppm). The cathode to be formed via evaporation of DMAc at 120 ◦ C for 2 h. The solid
and Li metal anode were separated by a Celgard 3501 separator. The coating atop slides was used as the precursor for laser irradiation. During
electrolyte used in this work was 1 M lithium bis- this process, the extremely high local temperature (>2500 oC) was
trifluoromethanesulfonyl imide (LiTFSI) dissolved in DME/DOL (1:1 in generated, [31,32] while the coatings were undergoing graphitization
volume) with 1.0 wt% LiNO3 as the additive, and the electrolyte/sulfur during the micro-explosive process, which facilitated the metal oxides to
ratio was controlled as about 10 μL mg− 1. The cathode slurry was pre­ undergo homogeneous nucleation process on the surface of the laser-
pared by mixing the LNPCs/S or LNPCs-ZnO/S cathode, super P, and induced porous carbons, and consequently forming transitional metal
polyvinylidene fluoride (PVDF) binder with a mass ratio of 8:1:1 in N- oxides/N-doped pore carbon hybrid frameworks (Fig. 1(b)). However,
methylpyrrolidone (NMP) and magnetic stirred for 2 h to form a ho­ the obtained hybrid frameworks have abundant internal voids, which
mogeneous dispersion. Then, the as-obtained slurry was uniformly cast can effectively encapsulate more active sulfur and absorb/store the
onto carbon-coated aluminum (Al) foil through using a doctor blade and electrolyte. To fabricate the LNPCs-ZnO/80S cathodes, a facile melt-
dried under vacuum oven at 60 ◦ C overnight. Subsequently, the ob­ diffusion method was used to infuse the sulfur powder into an internal
tained electrode foil was punched into disks with a diameter of 12 mm, void. As a result, the long-chain sulfur particles were uniformly
where the areal sulfur mass loading is around 2.4–6.7 mg cm− 2. Cyclic dispersed and anchoring on the surface of laser-induced pore carbons.
voltammetry (CV) plots and electrochemical impedance spectroscopy Based on the LNPCs-ZnO/80S cathodes, the Li-S battery was facilely
(EIS) were conducted on a CHI-760E electrochemical workstation assembled (Fig. 1(c)). For comparison, LNPCs hybrid frameworks
(Shanghai Chenhua Instrument Co. Ltd., China). CV plots were per­ without Zn moieties were also constructed with a similar method.
formed at a potential range of 1.7–2.8 V with the scan rate from 0.05 to To clarify the morphologies and structure features of LNPCs and
1 mV s− 1. EIS curves were implemented under open circuit potential LNPCs-ZnO hybrid frameworks, the FESEM and HRTEM images were
with a frequency ranging from 105 to 10-2 Hz at 5 mV amplitude. The collected. As shown in Fig. 2(a) and Figure S1 (Supporting Information),
galvanostatic charge–discharge (GCD) and rate performance were car­ the as-constructed LNPCs-ZnO hybrid frameworks were combined with
ried out using Land Battery Measurement System with the potential an inter-connected porous network by laser induction micro-explosion
from 1.6 to 2.8 V at room temperature. Shuttle current tests were con­ techniques under focusing conditions using PAA, glycine, and Zn
ducted similar to the previous work based on S. R. Narayanan method. (NO3)2 as the precursor, and the porous network with diameters of <
[87,88] LiNO3-free electrolyte was chosen to avoid Lithium anode 100 nm. In contrast, when only using PAA and glycine as the precursor,
passivation. For the measurement, the sulfur cathodes were fully acti­ the structure of LNPCs hybrid frameworks has changed significantly. As
vated under a current density of 0.05C. Then, the voltage of the cells was shown in Fig. 2(d), the LNPCs hybrid frameworks have a more homo­
charged to 2.7 V and leave the cell under open circuit potential for 5 geneous distribution of nanopores and macropores distributed
min. Subsequently, the test was switched to potentiostatic mode for 60 throughout the composites. Being consistent with the FESEM results, the
mins to offset the drop voltage at voltage stages of 2.6 V until steady- TEM image also revealed that ZnO nanoparticles with 5–10 nm were
state value reached. This steady-state current is the value of the shut­ uniformly anchored on the surfaces of LNPCs. Moreover, the HRTEM
tle current. The redox current tests of symmetric batteries and Li2S indicated that the ZnO nanoparticles were well crystallized with a
nucleation tests are shown in Supporting Information. marked lattice fringe spacing of ~ 0.26 nm, corresponding to the (0 0 2)
lattice plane (JCPDS no.79–0207) (Figure S1(d), Supporting Informa­
2.7. Theoretical calculation tion). [70] Fig. 2(e) displays the TEM images of the LNPCs hybrid
frameworks for comparison. The TEM image demonstrates that the
ZnO crystal structure with lattice parameters of a = b = 3.249 Å, c = fabricated LNPCs hybrid frameworks are composed of nanopores/mes­
5.205 Å was built. The studied (0 0 2) plane was then cleaved from the opores and hierarchical structures. At higher magnifications, it can be
unit cell. Further, the cleaved plane was increased to five atomic layers clearly observed that the nanopores/ mesopores are uniformly
and 3 × 3 super-cell, comprised of a total of 54 atoms, and a vacuum

3
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

Fig. 1. Schematic illustration of the construction of LNPCs-ZnO hybrid frameworks for Li-S batteries.

distributed on the LNPCs hybrid frameworks (Fig. 2(f)). Figure S2 (Supporting Information), the full survey XPS spectrum of
Raman spectroscopy of the LNPCs and LNPCs-ZnO hybrid frame­ LNPCs-ZnO hybrid frameworks reveals the presence of C 1 s (285.2 eV),
works was carried out, as shown in Fig. 2(g). The three typical peaks [76] O 1 s (532.1 eV), [23] N 1 s (398.2 eV), [29] and Zn 2p (1021.2 eV
located at around 1350, 1580, and 2650 cm− 1 clearly correspond to the and 1045.1 eV). [77,78] The deconvoluted XPS spectra of LNPCs-ZnO
D band arising from sp3-type disordered carbon and lattice defects, G hybrid frameworks at C 1 s region (Figure S2(b), Supporting Informa­
band arising from sp2-type ordered graphitic carbon and 2D band tion) show four main peaks at 284.8, 285.4, 286.2, and 289.4 eV,
resulting from second-order zone-boundary phonons, respectively. [68] respectively, corresponding to C–C, C-N/C-O, C-N/C-O, and O-C = O.
In comparison with LNPCs, LNPCs-ZnO hybrid frameworks can be [76] The C–C bond reveals that the LNPCs-ZnO hybrids have abundant
formed after adding Zn(NO3)2 as a micro-explosive initiator into PAA carbon frameworks. In the meantime, oxygen-containing functional
and glycine solutions. For LNPCs-ZnO hybrid frameworks, characteristic groups may play an important role in capturing polysulfide. The high-
peaks at 318 and 437 cm− 1 are ascribing to the bending vibration of Zn- resolution O 1 s spectrum in Figure S2(c) (Supporting Information)
O, suggesting the formation of ZnO nanoparticles. [71] Fig. 2(h) shows can be deconvoluted into four main peaks of Zn-O (530.9 eV), [77] C-O
the XRD patterns of LNPCs and LNPCs-ZnO hybrid frameworks. Both of (531.4 eV), [23] C = O (532.7 eV), [23] and O-C = O (533.8 eV). [23]
the LNPCs and LNPCs-ZnO hybrids show the obvious two broad pro­ Figure S2(d) (Supporting Information) shows the N 1 s spectra, and three
nounced diffraction peaks at 2θ ≈ 26.3◦ , and 43.2◦ , which can agree well prominent peaks can be distinguished at 398.3, 399.3, and 401.3 eV,
with the (0 0 2) and (0 0 4) diffraction peaks of graphitic structure of which correspond to Pyridinic N, Pyrrolic N, and Graphitic N, respec­
graphene (JCPDS 75–1621), respectively. [68,72,73] For LNPCs-ZnO tively. [29,34] Graphitic N can benefit for improving the conductivity of
hybrid frameworks, apart from the broad peak of LNPCs, sharp peaks the LNPCs-ZnO hybrid frameworks. [29,79] Pyrrolic N and Pyridinic N
at 2θ ≈ 31.7, 34.4, 36.2, 47.4, 56.4, 62.6, and 67.6◦ are observed, cor­ can bring about a large amount of functional and active site, and provide
responding to the (1 0 0), (0 0 2), (1 0 1), (1 0 2), (1 1 0), (1 0 3), and (1 1 2) multiple channels which further facilitate the electrolyte-ion diffusion.
planes of ZnO (JCPDS no.79–0207). [74,75] These results further [34] It’s suggested that the nitrogen-containing moieties can help
demonstrate the existence of ZnO nanoparticles in LNPCs-ZnO hybrid improving the electrochemical performance. [68] The Zn state spectrum
frameworks. Moreover, the chemical states and surface compositions of of LNPCs-ZnO hybrid frameworks shows two prominent peaks at 1021.2
the elements in the LNPCs and LNPCs-ZnO hybrids fabricated with and eV and 1045.1 eV (Figure S2(d), Supporting Information), correspond­
without addition of Zn(NO3)2 were analyzed by XPS. As depicted in ing to Zn 2p1/2 and Zn 2p3/2, respectively. [58,78] All these results

4
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

Fig. 2. Morphology and structure of LNPCs-ZnO and LNPCs hybrid frameworks. (a) FESEM images of LNPCs-ZnO with porous structure. (b) TEM image showing the
ZnO nanoparticles (dark color) and the background porous carbon-based materials. (c) HRTEM image of few-layer graphene with different interlayer distances ZnO
with 0.26 nm interlayer distance, respectively. (d) FESEM images of LNPCs with porous structure. (e) TEM image showing the nanopores/mesopores and hierarchical
structure of carbon-based materials. (f) HRTEM image of porous carbon with different pore diameters. (g) Raman spectroscopy of LNPCs and LNPCs-ZnO sample
showing the coexistence of nitrogen-doped graphene, ZnO nanoparticles, in comparison to the LNPCs from glycine. (h) XRD pattern of LNPCs and LNPCs-ZnO hybrids
showing characteristic peaks of nitrogen-doped graphene, and ZnO. (i) Nitrogen adsorption and desorption isotherms of LNPCs-ZnO hybrid frameworks.

strongly confirm that the unique structure of LNPCs-ZnO hybrid structure as the sulfur host, firstly the sulfur can be securely infiltrated
frameworks has been formed via this simple method. To further inves­ into the LNPCs-ZnO hybrid frameworks through a typical molten
tigate the porous structure of LNPCs and LNPCs-ZnO hybrid frameworks, method, resulting in substantial sulfur loading of 71.3–87.6 wt% in the
nitrogen adsorption–desorption isotherms and corresponding pore-size LNPCs-ZnO hybrid frameworks (Fig. 3(a)). In order to confirm that the
distribution are shown in Fig. 2(i) and Figure S4 (Supporting Informa­ sulfur powder was successfully diffused into the mesopores and micro­
tion). The LNPCs-ZnO hybrid framework possesses the largest specific pores of LNPCs-ZnO and LNPCs matrix, the XRD characterization was
surface area of 704.2 m2 g− 1 (with pore sizes below 2 nm), whereas carried out to measure pristine sulfur, LNPCs-ZnO/80S and LNPCs/80S
LNPCs has a value of 879.6 m2/g (with pore sizes <15 nm) (Figure S4, hybrids, respectively. As shown in Fig. 3(b), the additional sharp char­
Supporting Information). The huge specific surface area of LNPCs-ZnO acteristic peaks of LNPCs-ZnO/80S and LNPCs hybrids are well indexed
hybrids provides abundant reaction interface and facilitates the to the typical peaks of orthorhombic crystalline sulfur (JCPDS no.
dispersion of sulfur, thus further increasing the electrochemical capacity 08–0247), [81,82] demonstrating that the sulfur particles have been
of the cathode. [80] The pore volume resulting from interspace between encapsulated into the mesopores and micropores of LNPCs-ZnO and
adjacent hierarchical structures can mitigate the volume expansion, LNPCs matrix successfully. What’s more, compared to the diffraction
further improving the cycling stability. To estimate the weight content peaks of pristine crystalline sulfur, the peak intensity of the LNPCs-ZnO/
of ZnO in LNPCs-ZnO hybrid frameworks, the ICP-AES of the composite 80S and LNPCs/80S was drastically weakened. This indicates that the
was tested (Table S1, Supporting Information). Based on the ICP-AES crystalline sulfur formation has been suppressed efficiently due to the
result, the weight fraction of ZnO was about 4.33%. long-chain of S8 and most of the sulfur particles have been successfully
To investigate the performance of Li-S battery with the unique encapsulated within the pores of LNPCs-ZnO hybrids. To further

5
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

Fig. 3. (a) TGA curves of LNPCs-ZnO/70S, LNPCs-ZnO/80S, and LNPCs-ZnO/90S in N2 atmosphere. (b) XRD pattern of pristine sulfur, LNPCs-ZnO/80S and LNPCs/
80S hybrids. (c) XPS survey spectrum of LNPCs-ZnO/80S and LNPCs/80S hybrids. d-f) High-resolution XPS spectra of LNPCs-ZnO/80S with emphasis on C 1 s (d), Zn
2p (e) and S 2p (f) regions. (g) HAADF image of LNPCs-ZnO/80S hybrids. The corresponding EDX mapping of C (h), O (i), N (j), Zn (k), and S (l).

examine the surface compositions of the LNPCs-ZnO/80S hybrids, XPS LNPCs-ZnO/80S and corresponding elemental mapping images of C, O,
spectrum was performed (Fig. 3(c)). For comparison, XPS data of LNPCs N, Zn, and S are illustrated in Fig. 3(g-l) exhibiting that all the elements
were also collected (Figure S4, Supporting Information). The XPS are uniformly distributed in the interest area of composite indicating
spectra of LNPCs/80S and LNPCs-ZnO/80S hybrids in Fig. 3(c) and that the LNPCs-ZnO hybrid had successfully accommodated sulfur.
Figure S6 (Supporting Information) confirm the presence of C, N, O, Zn, To evaluate the electrochemical performance of the Li-S batteries
and S elements. The C 1 s XPS spectrum in Fig. 3(d) reveals the coex­ based on the LNPCs-ZnO/80S as the cathode, standard coin cells were
istence of four major peaks at 284.7, 285.7, 286.7, and 289.4 eV, which assembled and studied. LNPCs/80S cathode was also prepared for
can be attributed to C–C/C–H, C-N/C-O/C-S, C-N/C-O, and O = C-O, comparison. The initial five cyclic voltammetry (CV) curves of LNPCs-
respectively. [26,29] The characteristic peaks of Zn 2p1/2 and Zn 2p3/2 ZnO/80S cathode shown in Fig. 4(a) exhibit no obvious peak shifts
are observed at 1045.9 and 1022.3 eV in high-resolution Zn 2p spectra and peak current changes, demonstrating the excellent reversibility and
(Fig. 3(e)), which can be attributed to Zn2+ in ZnO. [58,78] Intriguingly, low-capacity decay of such electrochemical system with LNPCs/ZnO.
compared to those of the LNPCs-ZnO hybrids, the Zn 2p peaks in the The two representative reduction peaks locate at ~ 2.3 V and ~ 2.08 V,
LNPCs-ZnO/80S hybrids show a slight shift to higher binding energy, which are corresponded to the electrochemical transformation from S8
suggesting strong interactions between ZnO and sulfur. [58] In accor­ to long-chain polysulfides and then further reduction to insoluble short-
dance with the above results, two prominent peaks centered at 165.1 chain Li2S2 and Li2S, respectively, while the two oxidation peaks around
and 163.9 eV, which are assigned to the S 2p1/2 and S 2p3/2 for S or S8 in 2.38 V and 2.34 V versus Li/Li+ are due to the reverse oxidation process
the composite of LNPCs-ZnO/80S (Fig. 3(f)). [83] Moreover, two weaker from Li2S/Li2S2 to Li2S8 and then to S8. In contrast, the initial five CV
peaks appear at 161.7 and 168.7 eV, verifying that S2-/S2- 2 and S-O profiles of the Li-S battery with the LNPCs/80S cathode (Figure S8
bonding in sulfate, respectively. [28] The binding energy of the S 2p3/2 (Supporting Information)) indicate that the oxidation/reduction peaks
peak is lower than that of elemental sulfur (164.0 eV), further con­ are deformed and integral areas decreased, which suggests a poor
firming the possible existence of Zn-S interactions. [78] Figure S7 cycling stability and sluggish kinetic process. Figure S9(a) (Supporting
(Supporting Information) shows the high-resolution N 1 s spectra. Three Information) shows the initial curves of LNPCs-ZnO/80S and LNPCs/
prominent peaks correspond, respectively, to the Pyridinic N, Pyrrolic N, 80S cathodes, it is obvious to note that the reduction peaks of the LNPCs-
and N-oxide, which can be distinguished at 398.3, 400.2, and 402.7 eV. ZnO/80S cathode are positively shifted slightly and the oxidation peaks
[44,84] It is expected that the coexistence of N-doped heteroatoms and shift to the negative position, suggesting that ZnO NPs can improve the
ZnO NPs in the LNPCs-ZnO/80S hybrid can not only facilitate electro­ polysulfide redox kinetics and cell polarization decreases. [77] Figure S9
chemical reaction but also provide more efficient electroactive sites for (b) (Supporting Information) compares the first cycle charge–discharge
trapping the intermediates Li2Sx through strong chemical adsorption, voltage curves of the Li-S batteries with LNPCs/80S and LNPCs-ZnO/80S
while inhibiting the shuttling effect. [84] The typical TEM image of cathodes at 0.5C rate within a voltage window of 1.7–2.8 V versus Li+/

6
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

Fig. 4. Electrochemical performance of the LNPCs-ZnO/80S cathode. a) CV curves of the LNPCs-ZnO/80S at a scan rate of 0.1 mV s− 1. b) Galvanostatic char­
ge–discharge profiles of LNPCs-ZnO/80S cathodes at 1st,20th, 100th, 500th, and 1000th cycles. c) The rate performances of the Li-S batteries based on the LNPCs-
ZnO/80S or LNPCs/80S cathodes at various C-rates. d) Galvanostatic charge–discharge profiles of LNPCs-ZnO/80S cathodes at 0.2, 0.5, 1, 2, and 3C.e) Cycling
performance of the Li-S batteries with LNPCs-ZnO/80S or LNPCs/80S cathodes at 0.5C for 1000 cycles and corresponding coulombic efficiency.

Li. The LNPCs-ZnO/80S cathode shows two typical discharge plateaus at cathodes increased as the current density increased, and the discharge
around 2.3 and 2.1 V, and two obvious charge plateaus at around 2.38 capacities of only 328.5 mAh g− 1 was remained for LNPCs/80S cathode
and 2.34 V, which are well consistent with the result of CV profiles. From at 3C. The corresponding GCD profiles of the LNPCs-ZnO/80S and
Figure S9(b), it is obvious that the initial specific capacity of the LNPCs- LNPCs/80S cathode at different current densities are shown in Fig. 4(d)
ZnO/80S cathode (1184.9 mAh g− 1 at 0.5C) is much higher than that of and Figure S11 (Supporting Information). Although the polarization
the LNPCs/80S cathode (840.6 mAh g− 1 at 0.5C). This is a representa­ increased at high charge/discharge current densities, the profiles still
tive signal that polysulfide has been confined in the cathode porous showed two discharge plateaus and remained favorable reversibility. To
zone, polysulfides dissolution is efficiently restrained and further further verify the ZnO NPs on the improvement of the electrochemical
significantly inhibiting polysulfides shuttling effect. Fig. 4(b) shows the kinetics process during cycling, the electrochemical impedance spec­
GCD curves of the LNPCs-ZnO cathode with a sulfur content of 80% and troscopy (EIS) before/after cycling is tested (Figure S10, Supporting
at current densities of 0.5C (1C = 1675 mAh g− 1) after the 1st, 20th, Information). As shown in Figure S12 (Supporting Information), the
100th, 500th, and 1000th cycles. Two typical discharge plateaus can be LNPCs-ZnO/80S cathode before cycling (20.57 Ω.cm2) exhibits higher
clearly observed, corresponding to a typical multistep change of S8. The electrolyte resistance (Rs) compared with LNPCs/80S cathode (18.32
discharge voltage plateau, at 2.35 V, implies the reduction of S8 to Li2S8; Ω.cm− 2), which is mainly caused by the low electronic conductive ZnO
the corresponding discharge capacity (QH) is often used to assess the NPs that increased the resistance of the whole electrode. However, it can
polysulfides trapping efficiency and the degree of polysulfides genera­ be seen that the charge transfer resistance (Rct) value of LNPCs-ZnO/80S
tion. [85] Importantly, the QH in the 500 cycles was 79.1% of that in the cathode is 58.4 Ω, which are obviously lower than that of LNPCs/80S
initial cycle, indicating that the LNPCs-ZnO hybrid frameworks can be cathode (74.3 Ω). More importantly, Rs of the LNPCs-ZnO/80S (62 Ω)
suppressed the active sulfur loss efficiently. And the lower discharge after 1000th cycling is much smaller than that of LNPCs/80S (89.7 Ω),
plateau, at 2.05 V, indicates the transformation from polysulfides to suggesting that the ZnO NPs facilitates Li-ion diffusion between elec­
Li2S2/Li2S, which contributes to the main discharge capacity (70%). trolyte and electrode surface. In the low-frequency region, the slope of
Notably, the discharge plateaus of LNPCs-ZnO cathode display longer the line reflected ion diffusion rates in the electrode materials. Before
and flatter with a higher capacity and a lower polarization (△V = 0.23 cycling, the slope lines of LNPCs-ZnO/80S and LNPCs/80S show similar
V) than LNPCs (△V = 0.31 V) cathode (Figure S10 (Supporting Infor­ in Figure S12 (Supporting Information). After 1000th cycling, the slope
mation)), and only 29.5% fading is observed within 1000 cycles, sug­ of the LNPCs/80S cathode decreased, suggesting a lower ion diffusion
gesting a high polysulfides trapping ability of the ZnO NPs and LNPCs, coefficient of the hybrid frameworks. In contrast, LNPCs-ZnO/80S
and improved electrochemical reaction kinetic. The rate capabilities of cathode still has a big slope. In short, the outstanding performance of
the LNPCs-ZnO/80S and LNPCs/80S cathodes were further investigated LNPCs-ZnO/80S cathode are originated from the high conductivity of
under stepwise current densities from 0.2 to 3C (Fig. 4(c)). The LNPCs- 3D hybrid frameworks, polar ZnO NPs and N-doped porous carbons
ZnO/80S cathode delivered average reversible capacities of 1224.2, framework structures, which can efficiently trap the polysulfides,
1093.5, 960.6, 851, and 716.1 mAh g− 1 at current densities of 0.2, 0.5, further inhibiting shuttling effects.
1, 2 and 3C, respectively. Furthermore, after the high current densities, The cycling stabilities of the Li-S battery based on LNPCs-ZnO/80S
the LNPCs-ZnO/80S cathode exhibiting excellent capacity reversibility and LNPCs/80S cathode at 0.5C with the sulfur loading of 3.4 mg
and the discharge capacity recovers to 1169.1 mAh g− 1 at 0.2C. In cm− 2, are shown in Fig. 4(e). The initial discharge capacity of the Li-S
contrast, the rate performance of LNPCs/80S cathode is inferior to that battery with LNPCs-ZnO/S cathode is as high as 1155.8 mAh g− 1 (cor­
of LNPCs-ZnO/80S cathode, the capacity gaps between these two responding to areal capacity of 3.07 mAh cm− 2), which slowly decrease

7
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

to 760.7 mAh g− 1 (areal capacity: 2.11 mAh cm− 2) after 1000 cycles excellent long-term cycling performance (after 1000 cycles at 0.5C).
with an average capacity decay rate of 0. 31% per cycle and high- Compared with the cycled LNPCs/S electrode (Figure S13(a) and (b))
capacity retention of 70.5%, and with a superior Coulombic efficiency with cracks covered on the surface, the morphologies of the cycled
of nearly 100%. In striking contrast, the Li-S battery with LNPCs/S LNPCs-ZnO/S electrodes (Figure S13(c) and (d)) display smooth surface
cathode delivers an initial discharge capacity of 841.6 mAh g− 1 (areal and still maintains its original 2D hierarchical porous structure without
capacity: 2.26 mAh cm− 2), and drastically falls to 393.3 mAh g− 1 (areal any obvious pulverization or size variation as cycled before. Compared
capacity: 1.06 mAh cm− 2) after 1000 cycles with a Coulombic efficiency to the sulfur cathode, the introduction of ZnO NPs has a more profound
as high as 104.3% and with a high-capacity decay rate of 0.54% per influence on the lithium anode (Figure S14). It is clearly revealed that
cycle. It ascertains that LNPCs/S as a cathode show inferior cycle sta­ the lithium anode based on the LNPCs-ZnO/80S cathode (Figure S14(c)
bility, and the poor electrochemical performance can be contributed to and (d)) shows no cracks or dendrites but a smooth surface can be
few polar functional groups on the surface of LNPCs hybrid frameworks, observed. In sharp contrast, lithium anodes display heavy pitting
which cannot promote LiPS trapping efficiently, and hence the conver­ corrosion and obvious dendrites on the surface of the LNPCs/80S based
sion reaction of LiPS is sluggish. On the contrary, due to the stable and anode (Figure S14(a) and (b)). In brief, even after long cycling at high
high conductivity of LNPCs-ZnO/S frameworks and strong chemical current density, the lithium anode surface does not unveil a seriously
interaction between LiPSs and ZnO NPs, this unique structure demon­ damaged surface, demonstrating that the novel LNPCs-ZnO/80S cathode
strates that the polar ZnO NPs on the surface of high conductive N-doped can effectively block the notorious polysulfide shuttling to alleviate
porous carbons has suitable functional sites to effectively inhibit the lithium anode surface corrosion and further mitigate the anode cracks
shuttling effect. In order to further demonstrate the superior capture problem, thus benefiting to the superior long-term cycle stability and
mechanism of such electrochemical performance improvement, LNPCs- high sulfur utilization.
ZnO/S and LNPCs/S cells are disassembled. Post-cycling SEM To further investigate the chemical conversion of LiPS on the surface
(Figure S13 and Figure S14, Supporting Information) characterization of of LNPCs-ZnO and LNPCs, XPS spectra (Figure S15, Supporting Infor­
cathodes and lithium anodes provides additional evidence to verify the mation) were performed after 1000th charge/discharge process at 0.5C.

Fig. 5. (a) Cycling performance and the corresponding Coulombic efficiency of the Li-S battery based on the LNPCs-ZnO/S cathode at 1and 2C with a sulfur loadings
3.4 mg cm− 2. (b) Cycling performance and the corresponding coulombic efficiency of the Li-S battery based on the LNPCs-ZnO/S cathode at 0.2C with 3.4, 5.3, and
6.7 mg cm− 2 sulfur loadings, respectively. (c) Charge and discharge GITT curves at 0.2C. (d) CV curves of the symmetrical cells with the LNPCs and LNPCs-ZnO
electrodes in electrolytes with and without 0.5 M Li2S6 at a scan rate of 5 mV s− 1 (voltage range: –0.8 to 0.8 V). Fitting of current vs. time curve on different
surfaces: (e) LNPCs, and (f) LNPCs-ZnO. The cells were discharged galvanostatically at 0.1 mA to 2.06 V and kept potentiostatically at 2.05 V until the current was
below 10-5 A. (g) Schematic representation of flexible Li-S pouch cell employing a LNPCs-ZnO/S cathode. (h) Cycling performance of pouch cell at different folding
angles and the corresponding pictures of LEDs lighten up by pouch cell.

8
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

At the fully-discharged state of 1.7 V (Figure S15(c)), the obvious ratio is significantly lower than other ratios, only 141 Wh/kg can be
characteristic peak located at 161.4 and 162.5 eV, which are attributed achieved (Figure S18(b), Supporting Information).
to the fully transformation of LiPSs into lithium sulfide (Li2S2/Li2S). To investigate the reaction kinetics of the LPSs transformation
Furthermore, it is obviously noted that the peak of the lithium sulfide behavior for the LNPCs-ZnO/80S and LNPCs/80S cathodes, CV mea­
species gradually decreases after the subsequent fully-charged state of surements were carried out between 1.7 and 2.8 V at various scan rate
2.8 V (Figure S15(d)). Besides, the peak of the S-S bond obviously in­ from 0.05 to 1 mV s− 1 and the relationship of peak current versus (v/
creases. Intriguingly, compared to those of the LNPCs-ZnO hybrids, the s)0.5 are displayed in Figure S20 (Supporting Information). As shown in
peak of the lithium sulfide in the LNPCs/80S hybrids show incomplete Figure S20 (a) and (d), LNPCs-ZnO/80S and LNPCs/80S cathodes show
conversion (Figure S15(b)), suggesting the presence of ZnO can effec­ two representative reduction peaks and two oxidation peaks for all scan
tively promote the reversible oxidation of lithium sulfide to LiPSs. rates. Notably, because of the diffusion process of LiPS, a linear rela­
For the LNPCs-ZnO/80S cathode, the cycling performances of Li-S tionship between the peak current and the square root of scan rate can
batteries at a larger current density of 1 and 2C were further detected, be determined by diffusion coefficient of Li ions (DLi+) according to the
respectively (Fig. 5(a)). Clearly, the LNPCs-ZnO/80S cathodes show classical Randles-Sevcik equation:
high reversible specific capacities of 952.9 mAh g− 1 (1C) and 858.9 mAh
g− 1 (2C), which is 25.9% and 39.3% higher than that of LNPCs/80S Ip = 2.69*105n1.5AD0.5
Li CLiv
0.5
(1)
cathode (705.3 mAh g− 1 at 1C and 521.6 mAh g− 1 at 2C), respectively. where Ip is the peak current, n is the charge transfer number, A is the
The corresponding areal capacities are as high as 2.81 (1C) and 2.53 area of active electrode, DLi is the lithium ion diffusion coefficient, CLi is
mAh cm− 2 (2C) for the LNPCs-ZnO/80S cathode, which are much higher the concentration of lithium ions in the electrolyte, and v is the scan rate.
than those of the LNPCs/80S cathode (2.07 mAh cm− 2 at 1C and 1.54 Therefore, the slope of curves (Ip/V0.5) can reflect the lithium-ion
mAh cm− 2 at 2C) (Figure S16, Supporting Information). The unique diffusion rate, which provides critical information about the kinetics
structure of the LNPCs-ZnO hybrids makes the battery to achieve better of polysulfides conversion, because n, A, and CLi in the battery are un­
cycling stability with a discharge capacity of 725.9 mAh g− 1 at 1C (2.14 changed. Clearly, the LNPCs-ZnO/80S cathode exhibits larger slopes
mAh cm− 2) and 525 mAh g− 1 at 2C (1.54 mAh cm− 2) after 500 cycles, (bIA = 8.07, bIC1 = -4.57, bIC2 = -5.94) as compared with the LNPCs/80S
exceeding those of 310.9 mAh g− 1 at 1C (0.91 mAh cm− 2) and 116.6 cathode (bIA = 6.92, bIC1 = -3.40, bIC2 = -5.75). The resulting DLi in
mAh g− 1 at 2C (0.34 mAh cm− 2) in LNPCs/80S cathode, respectively. peaks IA, IC1, and IC2 are 4.56*10-8, 1.46*10-8, 2.47*10-8 cm2/s for
The Coulombic efficiencies of LNPCs-ZnO/80S cathodes at 1 and 2C LNPCs-ZnO/80S and 3.36*10-8, 8.10*10-9, 2.32*10-8 cm2/s for LNPCs/
both are near 100%. By comparison experiments, it’s found that the 80S, respectively. In addition, galvanostatic inter-mittent titration
excellent cycling stability is obtained by LNPCs-ZnO/80S cathodes, technique (GITT) was also conducted to determine the Li+ diffusion
which may be mainly ascribed to high-efficiency conductive networks coefficients. Fig. 5(c) shows the GITT profiles for both LNPCs and
and superior structural stability. The ZnO NPs anchoring on the surface LNPCs-ZnO. It shows an obvious lower polarization between the
of LNPCs provides more active sites for trapping polysulfides and pro­ discharge and charging plateau for the LNPCs-ZnO, which should be
motes polysulfides fast diffusion. [58,77,78] Moreover, considering the related to the promotion of redox kinetics by ZnO NPs. Apparently, the
high sulfur loading and high reversible capacity of the LNPCs-ZnO/S Li+ diffusion coefficients (DLi+) of LNPCs-ZnO hybrid frameworks
cathode, the cycling stability of Li-S battery assembled with the sulfur (8.47*10-8 cm2/s) are larger than LNPCs (9.79*10-8 cm2/s). Those result
loading of 3.4, 5.3, and 6.7 mg cm− 2 at 0.5C is investigated. As shown in are consistent with the above CV curve calculation results. Above all, it
Fig. 5(b), with the sulfur loading of 3.4, 5.3, and 6.7 mg cm− 2 at 0.2C, clearly indicates that the faster diffusion process of LiPS is mainly
the Li-S battery exhibits initial areal capacities of 2.93, 5.01, and 6.36 related to the introduction of ZnO NPs anchoring on the LNPCs frame­
mAh cm− 2, respectively. After 200 cycles, the reversible areal capacities works, which can strong adsorb LiPSs and facilitate the sulfur redox
maintain at 2.45, 3.85, and 4.71 mAh cm− 2, achieving a capacity reactions for further enhancing the Li storage performance.
retention of 74–84%. Even when the sulfur loading increases up to 6.7 In addition, the power-law formula i = avb was used to analyze the
mg cm− 2, the high reversible areal capacity of 4.71 mAh cm− 2 is still reaction kinetics. The values of b can be easily calculated from the slope
retained after 200 cycles. Such a high areal capacity of the LNPCs-ZnO/S of the linear plots of log(sweep rate) - log(peak current). In Figure S20
cathode can be contributed to its unique structures, in which the 3D (c), the fitted b-values of the LNPCs-ZnO/80S are determined to be 0.62,
conductive networks facilitates rapid electronic/ionic transport, and the 0.60, and 0.58, respectively, but the LNPCs/80S (Figure S20(f)) shows
high specific surface area offers multiple void spaces, which provides lower b values (0.47, 0.53, and 0.52, respectively) than those of the
more storage space for sulfur and further beneficial to improve sulfur LNPCs-ZnO/80S. Those results suggest that the reversible conversion
utilization. To the best of our knowledge, even under high areal sulfur kinetics of polysulfides can be significantly enhanced by LNPCs-ZnO/
loading and high-rate charging-discharging state, this high electro­ 80S electrode. Further research reveal that the b values corresponding
chemical performance of LNPCs-ZnO/S cathodes was superior to many to IC1 of LNPCs-ZnO/80S and LNPCs/80S are significantly greater than
previously reported TMO-based cathodes materials (Tables S3 and S4 that of IC2, which is indicated that the conversion kinetics of S8 → Li2S6/
Supporting Information), highlighting the great potential of as-prepared Li2S4 (solid–liquid conversion) are faster than Li2S6/Li2S4 → Li2S2/Li2S
LNPCs-ZnO hybrid frameworks for next-generation high-performance (liquid–solid conversion). Importantly, the sluggish Li reaction kinetics
Li-S batteries. Because the electrode material has a strong trapping on the surface of the LNPCs frameworks can be enhanced by ZnO NPs.
ability to polysulfides, it can effectively suppress polysulfide shuttles. To further demonstrate the enhanced redox kinetics catalyzed by
According to previous literature reports, shuttle current can be used as ZnO, redox current in combination with Li2S nucleation experiments
one of the important parameters to evaluate the degree of shuttle. were carried out. The redox current curves were collected by the CV
Figure S17 compares the shuttle behavior of LNPCs/80S and LNPCs- measurements of symmetric cells with loading LNPCs-ZnO (or LNPCs)
ZnO/80S cathodes. As can be seen from FigureS17, the LNPCs-ZnO/ on steel mesh simultaneously serving as the working and counter elec­
80S cathode shows a smaller shuttle current compared with the trodes and Li2S6 solution (0.5 M) served as the electrolyte at a scan rate
LNPCs/80S cathode, indicating that the LNPCs-ZnO hybrid frameworks of 50 mV/s in the voltage range from − 0.8 to 0.8 V. As shown in Fig. 5
have a strong trapping ability to suppress polysulfide shuttling effect. (d), LNPCs-ZnO electrode presents two pairs of pronounced redox peaks
Figure S18 compares the cycling performance of the Li-S battery based located at − 0.42, 0.03, 0.42, and − 0.03 V, demonstrating the reversible
on the LNPCs-ZnO/S cathode at 0.5C with 5, 10, and 15 μL/mg E/S ratio, conversion of the Li2S6 in electrochemical process. In contrast, the
respectively. As can be seen from Figure S18, the E/S ratio of 15 μL/mg symmetric cell with LNPCs electrode only deliver the unobvious redox
shows a better cycling stability compared with the 5 and 10 μL/mg. peaks with the lower current response than the case of the LNPCs-ZnO
However, the energy density of lithium-sulfur batteries under this N/P

9
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

electrode. It further implies that the outstanding catalytic ability of ZnO ZnO based Li-S pouch cells could be attributed to the excellent cycling
toward LiPS redox. In parallel, Li2S precipitation experiments are driven stability as well as the effective suppression of polysulfide shuttling,
by loading LNPCs-ZnO and LNPCs on carbon fiber paper (CFP), highlighting their practical use in flexible energy systems.
respectively. It is observed that a sharp nucleation peak appears earlier To shed light on the catalysis process from the perspective of mo­
in the LNPCs-ZnO cell (Fig. 5(f)) than in the LNPCs cell (Fig. 5(e)). More lecular or atomic scales, the density functional theory (DFT) calculations
importantly, a large capacity of 175 mAh gs− 1 (Fig. 5(f)) indicates that were conducted to elucidate how LiPSs were activated during catalysis
LNPCs-ZnO can induce more nucleation of Li2S than LNPCs (116 mAh and the binding energy between ZnO NPs (or LNPCs) and long-chain
gs− 1), suggesting the accelerated Li2S precipitation in the presence of Li2Sn (n = 4, 6, 8)/S8. As aforementioned, ZnO (0 0 2) facet (shown in
ZnO. Above all, ZnO NPs anchoring on the LNPCs frameworks can Figure S21(a)) was selected as the representative plane for modeling
efficiently address the LiPSs shuttle and enhance LiPS conversion ki­ because of its larger surface area and provided favorable adsorption
netics, which would be beneficial to modifying sulfur utilization and locations for Li2Sx species. The binding energies (Eb) of Li2Sn/S8 (mo­
showing excellent reversible capacity and superior rate capability of lecular structures are shown in Figure S21 (b)) with ZnO and LNPCs
LNPCs-ZnO/80S cathode-based Li-S battery. were implemented by a DMol3 Code in the framework of DFT. Fig. 6(a)
To further demonstrate the practical application of as-designed reveals the binding geometric configurations between ZnO and various
LNPCs-ZnO electrodes, a flexible Li-S was assembled as pouch cells sulfur species, with the corresponding Eb values marked. In the
(Fig. 5(g)). As expected, 41 red light diode (LED) array of “SJTU” letters adsorption structure (Fig. 6(a)), the most favorable binding sites of
were lightened up by pouch cells (Fig. 5(h)). To show the great me­ Li2Sn/S8 are S atoms which tend to bond with the Zn atoms on ZnO
chanical flexibility, different folding angles for pouch cells were tested. (0 0 2), in which the S-Zn distance is 2.93 Å. Clearly, the Eb values be­
As shown in Fig. 5(h), the pouch cells were subjected to bending tween the (0 0 2) facet of ZnO and S8, Li2S, Li2S2, Li2S4, Li2S6, and Li2S8
deformation and almost no change in brightness was observed even species are − 3.94, − 6.91, − 5.52, − 4.34, − 3.84 and − 5.21 eV, respec­
when bending at an angle of nearly 180◦ . Furthermore, the electro­ tively. As shown in Fig. 6(a), the Eb value for Li2S8 remarkably reaches
chemical performance was also monitored to show the cycle stability in − 5.21 eV, which is one of the best-reported values for ZnO-based sys­
various folding angles at 0.2C (Fig. 5(h)). The capacity loss was negli­ tems. When the molecular length of LiPSs further decreased, the corre­
gible even at the folding angles of 180◦ . Moreover, the areal capacity of sponding Eb values decrease to the lowest value of − 3.84 eV for Li2S6.
pouch cells decreased a little from initial 3.23 to 2.47 mAh cm− 2 after After that, the Eb values gradually increase to − 6.91 eV for Li2S, which
400 cycles. The favorable electrochemical performance of the LNPCs- could be interpreted by the strong interaction between the ZnO surface

Fig. 6. Theoretical calculation and adsorption test of LNPCs-ZnO with Li2Sn/S8. (a) The trends of binding energies for various polysulfides molecular with (0 0 2)
plane of ZnO. (b) UV–visible spectra of the Li2S6 solution after exposure to LNPCs, and LNPCs-ZnO, respectively. Insets are the digital photograph of the Li2S6 and
polysulfides trapping ability by LNPCs, and LNPCs-ZnO before and after one week. (c) Free Gibbs energy curves for the reduction of LiPSs on the LNPCs (black line)
and ZnO (red line). Insets are the optimized adsorption geometries of intermediate species on the LNPCs and ZnO. Li2S decomposition on LNPCs (d) and ZnO (f) and
Li2S decomposition path on ZnO. (g) Schematic of the capture and catalytic mechanism of the LiPSs on the surface of LNPCs-ZnO in electrochemical process.

10
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

and short-chain polysulfides owing to similar polarity. On the other enhance the cycle life and rate performance of the Li-S battery. Conse­
hand, the corresponding Eb values of S8, Li2S, Li2S2, Li2S4, Li2S6, and quently, the LNPCs-ZnO/S cathode with 80 wt% of sulfur (areal sulfur
Li2S8 species on LNPCs surface are − 0.43, − 1.32, − 1.27, − 0.90, − 0.98, loading of 3.4 mg cm− 2) delivers a high initial discharge capacity of
and − 1.07 eV, respectively (Figure S22, Supporting Information). Ac­ 1155.8 mAh g− 1 at 0.5C with a Coulombic efficiency of 99.7%.
cording to the above DFT results, a higher Eb values of ZnO to LiPSs Furthermore, the LNPCs-ZnO/80S cathode can operate at 0.5C over
indicates a stronger chemical anchoring of LiPSs. 1000 cycles with 70.5% capacity retention and shows superior rate
In order to further evaluate the adsorption ability of LiPSs on the performance up to 3C. Even when the areal sulfur loading reaches up to
surface of ZnO, the LiPSs adsorption test using LNPCs and LNPCs-ZnO 6.7 mg cm− 2, the LNPCs-ZnO/S cathode exhibits high initial areal ca­
hybrid frameworks with the 5 mmol Li2S6 in DOL/DME (1:1, v/v) are pacity of 6.36 mAh cm− 2 and excellent cycling stability at 0.2C. In
showing in Fig. 6(b). Interestingly, after a short static test of one day, addition, the adsorption test and DFT calculations demonstrate that the
LNPCs-ZnO almost decolorizes while the LNPCs sample in which still intrinsic interaction between LiPSs/S8 and ZnO NPs is the main factor to
retained yellow. Moreover, even after one week of the static adsorption inhibit the shuttle effect of LiPSs. This study will indeed provide a new
test, the color of the LNPCs-ZnO-containing solution changes to almost direction to construct various 3D porous hybrid frameworks to be
transparent compared to the LNPCs in which the color of Li2S6 is still applied in the next-generation high-performance wearable and portable
retained. These results indicated that the LNPCs-ZnO-based cathode has devices.
a strong absorbing capacity on polysulfides. To further verify the strong
Li2S6 adsorption ability of LNPCs and LNPCs-ZnO, the corresponding Declaration of Competing Interest
ultraviolet–visible (UV–vis) spectroscopy was also measured. As can be
seen from Fig. 6(b), the absorption peak of Li2S6 solution in the 410–420 The authors declare that they have no known competing financial
nm region are disappeared obviously after adding LNPCs-ZnO, and in interests or personal relationships that could have appeared to influence
contrast, obvious absorbance is seen for those exposed to LNPCs. As the work reported in this paper.
theoretical and experimental results substantially reveal that LNPCs-
ZnO is an ideal cathode material, ZnO NPs can be considered as a type Acknowledgements
of polar matrix and can provide strong chemical interaction with LiPS,
which can effectively confine LiPSs shuttling and further improve the The financial supports from the National Natural Science Foundation
rate and cycle performances of Li-S battery. of China (no. 61574091), China Postdoctoral Science Foundation
Fig. 5(c) shows the calculated Gibbs free energies of the stepwise (2020M681300), Shanghai Sailing Program (18YF1427800) and the
intermediates including S8, Li2S, Li2S2, Li2S4, Li2S6, and Li2S8 species on Foundation for SMC Excellent Young Teacher in Shanghai Jiao Tong
the LNPCs and ZnO during the polysulfide transformation reaction. The University were gratefully acknowledged. The analysis supports from
reduction from Li2S2 to Li2S has the largest positive Gibbs free energy in the Instrumental Analysis Center of Shanghai Jiao Tong University and
the whole discharging process, indicating that the rate-controlling step the Center for Advanced Electronic Materials and Devices of Shanghai
and the sluggish sulfur redox reaction in the Li-S battery. And the value Jiao Tong University were also acknowledged.
on ZnO (0.67 eV) is much lower than LNPCs (1.12 eV), suggesting that
the presence of ZnO on LNPCs enables the reduction from Li2S2 to Li2S is Appendix A. Supplementary data
thermodynamically more favorable than pristine LNPCs. In addition, the
Li2S oxidation process during the inverse charging process play an Supplementary data to this article can be found online at https://doi.
important role for studying reaction kinetics in electrochemical process. org/10.1016/j.cej.2021.129707.
As shown in Fig. 6(d) and (e), the calculated Li2S decomposition energy
barrier of ZnO surface (1.51 eV) is far lower than the case of LNPCs References
(2.49 eV). Such a low energy barrier confirms that the ZnO NPs on
LNPCs can efficiently catalyze the breaking of the Li-S bond and enhance [1] A. Manthiram, Y. Fu, S.H. Chung, C. Zu, Y.S. Su, Chem. Rev. 114 (2014)
11751–11787.
delithiation reaction kinetics during the electro-chemical process.
[2] Y.X. Yin, S. Xin, Y.G. Guo, L.J. Wan, Angew. Chem. Int. Ed. Engl. 52 (2013)
Benefiting from merits of the transition metal oxides in terms of 13186–13200.
effective confinement, catalysis and diffusion, the Li-S batteries assem­ [3] N. Jayaprakash, J. Shen, S.S. Moganty, A. Corona, L.A. Archer, Angew. Chem. Int.
bled with LNPCs-ZnO/S cathodes show excellent electrochemical per­ Ed. Engl. 50 (2011) 5904–5908.
[4] M. Zhao, B.Q. Li, X. Chen, J. Xie, H. Yuan, J.Q. Huang, Chem. 6 (2020) 3297–3311.
formance. As illustrated in Fig. 6(h), the polar ZnO NPs in LNPCs-ZnO [5] Y. Zhao, Y. Ye, F. Wu, Y. Li, L. Li, R. Chen, Adv. Mater. 31 (2019), e1806532.
electrode exhibits strong capture ability of soluble polysulfides, which [6] J. Liu, T. Qian, N. Xu, M. Wang, J. Zhou, X. Shen, C. Yan, Energy Storage Mater. 24
effective suppress polysulfide shuttling outside of cathode. Conse­ (2020) 265–271.
[7] M. Zhong, J. Guan, J. Sun, H. Guo, Z. Xiao, N. Zhou, Q. Gui, D. Gong, Electrochim.
quently, sulfur utilization during charge–discharge processes is Acta 299 (2019) 600–609.
improved and cycling stability is ultimately extended. Moreover, the [8] M. Zhao, X. Chen, X.Y. Li, B.Q. Li, J.Q. Huang, Adv. Mater. 32 (2021) 2007298.
catalytic effect of multifunctional ZnO NPs can promote fast conversion [9] X. Ji, K.T. Lee, L.F. Nazar, Nat. Mater. 8 (2009) 500–506.
[10] J. Zhang, H. Ye, Y. Yin, Y. Guo, J. Energy Chem. 23 (2014) 308–314.
from polysulfides to Li2S, high areal sulfur loading is able to undergo [11] M.Q. Zhao, Q. Zhang, G.L. Tian, J.Q. Huang, W. Zhu, F. Wei, ACS Nano 6 (2012)
complete redox reactions and regulatable Li2S deposition sites, which 10759.
facilitate the fast polysulfides conversion kinetics, high areal capacity, [12] H.J. Peng, J.Q. Huang, M.Q. Zhao, Q. Zhang, X.B. Cheng, X.Y. Liu, W.Z. Qian,
F. Wei, Adv. Funct. Mater. 24 (2014) 2772–2781.
and superior cycling stability of Li-S batteries.
[13] X.B. Cheng, J.Q. Huang, Q. Zhang, H.J. Peng, M.Q. Zhao, F. Wei, Nano Energy 4
(2014) 65–72.
4. Conclusion [14] G. Zheng, Y. Yang, J.J. Cha, S.S. Hong, Y. Cui, Nano Lett. 11 (2011) 4462–4467.
[15] R. Elazari, G. Salitra, A. Garsuch, A. Panchenko, D. Aurbach, Adv. Mater. 23 (2011)
5641–5655.
In summary, a laser-induced metal oxide/N-doped porous carbons [16] R. Singhal, S.-H. Chung, A. Manthiram, V. Kalra, J. Mater. Chem. A 3 (2015)
hybrid framework was firstly demonstrated for high-performance Li-S 4530–4538.
[17] H. Guang, E. Scott, L. Xiao, C. Marine, G.N. Arnd, F. Linda, ACS Nano 7 (2013)
batteries. As a cathode in Li-S batteries, the unique structure of LNPCs-
10920–10930.
ZnO hybrid frameworks not only provides adequate internal voids, but [18] Z. Li, J. Zhang, B. Guan, D. Wang, L.M. Liu, X.W. Lou, Nat. Commun. 7 (2016)
also greatly improves the conductivity of the cathode. In addition, these 13065–13072.
unique structures could efficiently suppress the shuttle effects due to the [19] H. Wang, Y. Yang, Y. Liang, J.T. Robinson, Y. Li, A. Jackson, Y. Cui, H. Dai, Nano
Lett. 11 (2011) 2644–2647.
superior trapping ability of LiPSs by ZnO NPs, which can endow ZnO [20] G. Zhou, S. Pei, L. Li, D.W. Wang, S. Wang, K. Huang, L.C. Yin, F. Li, H.M. Cheng,
NPs and its hybrid frameworks with an impressive ability to significantly Adv. Mater. 26 (2014) 625–631.

11
C. Zhou et al. Chemical Engineering Journal 421 (2021) 129707

[21] G. Zhou, L. Li, C. Ma, S. Wang, Y. Shi, N. Koratkar, W. Ren, F. Li, H.-M. Cheng, [54] Z. Sun, J. Zhang, L. Yin, G. Hu, R. Fang, H.M. Cheng, F. Li, Nat. Commun. 8 (2017)
Nano Energy 11 (2015) 356–365. 14627–14636.
[22] C. Huang, Y. Zhou, H. Shu, M. Chen, Q. Liang, S. Jiang, X. Li, T. Sun, M. Han, [55] N. Li, Y. Xie, S. Peng, X. Xiong, K. Han, J. Energy Chem. 42 (2020) 116–125.
Y. Zhou, J. Jian, X. Wang, Electrochim. Acta 329 (2020), 135135. [56] Z. Xiao, Z. Li, P. Li, X. Meng, R. Wang, ACS Nano 13 (2019) 3608–3617.
[23] W. Sun, Y. Xu, X. Chen, Y. Xu, F. Wu, Y. Wang, Chem. Eng. J. 383 (2020), 123111. [57] P. Zhao, Z. Zhang, H. He, Y. Yu, X. Li, W. Xie, Z. Yang, J. Cai, ChemSusChem 12
[24] G. Hu, C. Xu, Z. Sun, S. Wang, H.M. Cheng, F. Li, W. Ren, Adv. Mater. 28 (2016) (2019) 4866–4873.
1603–1609. [58] J. Zhang, P. Gu, J. Xu, H. Xue, H. Pang, Nanoscale 8 (2016) 18578–18595.
[25] Z. Wang, Y. Dong, H. Li, Z. Zhao, H.B. Wu, C. Hao, S. Liu, J. Qiu, X.W. Lou, Nat. [59] T. Zhao, Y. Ye, X. Peng, G. Divitini, H.-K. Kim, C.-Y. Lao, P.R. Coxon, K. Xi, Y. Liu,
Commun. 5 (2014) 5002. C. Ducati, R. Chen, R.V. Kumar, Adv. Funct. Mater. 26 (2016) 8418–8426.
[26] W. Jun, C. Bing, L. Qingqing, H. Ailin, L. Xiaoying, J. Qi, Scripta Mater. 177 (2020) [60] M. Yu, A. Wang, F. Tian, H. Song, Y. Wang, C. Li, J.D. Hong, G. Shi, Nanoscale 7
208–213. (2015) 5292–5298.
[27] Z. Jia, H. Zhang, Y. Yu, Y. Chen, J. Yan, X. Li, H. Zhang, J. Energy Chem. 43 (2020) [61] M. Zhao, B.Q. Li, H.J. Peng, H. Yuan, J.Y. Wei, J.Q. Huang, Angew. Chem. Int. Ed.
71–77. 59 (2020) 12636–12652.
[28] B. Gan, K. Tang, Y. Chen, D. Wang, N. Wang, W. Li, Y. Wang, H. Liu, G. Wang, [62] C. Lin, L. Qu, J. Li, Z. Cai, H. Liu, P. He, X. Xu, L. Mai, Nano Res. 12 (2018)
J. Energy Chem. 42 (2020) 174–179. 205–210.
[29] H. Chen, W.D. Dong, F.J. Xia, Y.J. Zhang, M. Yan, J.P. Song, W. Zou, Y. Liu, Z. [63] J. Lin, Z. Peng, Y. Liu, F. Ruiz-Zepeda, R. Ye, E.L. Samuel, M.J. Yacaman, B.
Y. Hu, J. Liu, Y. Li, H.E. Wang, L.H. Chen, B.L. Su, Chem. Eng. J. 381 (2020), I. Yakobson, J.M. Tour, Nat. Commun. 5 (2014) 5714–5726.
122746. [64] R. Ye, D.K. James, J.M. Tour, Adv. Mater. 31 (2019) 1803621.
[30] A. Bhargav, J. He, A. Gupta, A. Manthiram, Joule 4 (2020) 285–291. [65] Y. Chyan, R. Ye, Y. Li, S.P. Singh, C.J. Arnusch, J.M. Tour, ACS Nano 12 (2018)
[31] Q. Zhu, H. Deng, Q. Su, G. Du, Y. Yu, S. Ma, B. Xu, Electrochim. Acta 293 (2019) 2176–2183.
19–24. [66] X. Lu, G. Wu, Q. Xiong, H. Qin, W. Wang, F. Luo, Appl. Surf. Sci. 406 (2017)
[32] X. Wu, N. Liu, M. Wang, Y. Qiu, B. Guan, D. Tian, Z. Guo, L. Fan, N. Zhang, ACS 265–273.
Nano 13 (2019) 13109–13115. [67] F. Zhang, E. Alhajji, Y. Lei, N. Kurra, H.N. Alshareef, Adv. Energy Mater. 8 (2018)
[33] R. Wang, J. Yang, X. Chen, Y. Zhao, W. Zhao, G. Qian, S. Li, Y. Xiao, H. Chen, Y. Ye, 1800353.
G. Zhou, F. Pan, Adv. Energy Mater. 10 (2020) 1903550. [68] C. Zhou, K. Zhang, M. Hong, Y. Yang, N. Hu, Y. Su, L. Zhang, Y. Zhang, Chem. Eng.
[34] X. Wen, K. Xiang, Y. Zhu, L. Xiao, H. Liao, W. Chen, X. Chen, H. Chen, J. Alloys J. 385 (2020), 123720.
Compd. 815 (2020), 152350. [69] F. Clerici, M. Fontana, S. Bianco, M. Serrapede, F. Perrucci, S. Ferrero, E. Tresso,
[35] X. Wen, X. Lu, K. Xiang, L. Xiao, H. Liao, W. Chen, W. Zhou, H. Chen, J. Colloid A. Lamberti, A.C.S. Appl, Mater. Interfaces 8 (2016) 10459.
Interface Sci. 554 (2019) 711–721. [70] D.I. Son, B.W. Kwon, D.H. Park, W.S. Seo, Y. Yi, B. Angadi, C.L. Lee, W.K. Choi, Nat
[36] P. Vélez, M.L. Para, G.L. Luque, D. Barraco, E.P.M. Leiva, Electrochim. Acta 309 Nanotechnol. 7 (2012) 465–471.
(2019) 402–414. [71] M. Morsy, I.S. Yahia, H.Y. Zahran, F. Meng, M. Ibrahim, J. Electron. Mater. 48
[37] T. Yang, K. Liu, R. Ren, J. Zhang, X. Zheng, C. Wang, M. Chen, Chem. Eng. J. 381 (2019) 7328–7335.
(2020), 122685. [72] X. Zang, C. Shen, Y. Chu, B. Li, M. Wei, J. Zhong, M. Sanghadasa, L. Lin, Adv.
[38] J. Zhang, W. Ma, Z. Feng, F. Wu, D. Wei, B. Xi, S. Xiong, J. Energy Chem. 39 (2019) Mater. 30 (2018), e1800062.
54–60. [73] J. Cai, C. Lv, A. Watanabe, Nano Energy 30 (2016) 790–800.
[39] X. Chen, Y. Xu, F.H. Du, Y. Wang, Small Methods 3 (2019) 1900338. [74] M. Azarang, A. Shuhaimi, R. Yousefi, S.P. Jahromi, RSC Adv 5 (2015)
[40] X. Liu, J.Q. Huang, Q. Zhang, L. Mai, Adv. Mater. 29 (2017) 1601759. 21888–21896.
[41] X. Tao, J. Wang, C. Liu, H. Wang, H. Yao, G. Zheng, Z.W. Seh, Q. Cai, W. Li, [75] M. Fu, Y. Li, S. Wu, P. Lu, J. Liu, F. Dong, Appl. Surf. Sci. 258 (2011) 1587–1591.
G. Zhou, C. Zu, Y. Cui, Nat. Commun. 7 (2016) 11203. [76] L. Lu, J.T.M. De Hosson, Y. Pei, Carbon 144 (2019) 713–723.
[42] W.G. Lim, S. Kim, C. Jo, J. Lee, Angew. Chem. Int. Ed. 58 (2019) 18746–18757. [77] R. Yang, H. Du, Z. Lin, L. Yang, H. Zhu, H. Zhang, Z. Tang, X. Gui, Carbon 141
[43] Y. You, Y. Ye, M. Wei, W. Sun, Q. Tang, J. Zhang, X. Chen, H. Li, J. Xu, Chem. Eng. (2019) 258–265.
J. 355 (2019) 671–678. [78] N. Shi, B. Xi, Z. Feng, F. Wu, D. Wei, J. Liu, S. Xiong, J. Mater. Chem. A 7 (2019)
[44] W. Yang, W. Yang, L. Dong, X. Gao, G. Wang, G. Shao, J. Mater. Chem. A 7 (2019) 4009–4018.
13103–13112. [79] Q. Li, J. Guo, J. Zhao, C. Wang, F. Yan, Nanoscale 11 (2019) 647–655.
[45] Z.A. Ghazi, X. He, A.M. Khattak, N.A. Khan, B. Liang, A. Iqbal, J. Wang, H. Sin, [80] D. Cheng, P. Wu, J. Wang, X. Tang, T. An, H. Zhou, D. Zhang, T. Fan, Carbon 143
L. Li, Z. Tang, Adv. Mater. 29 (2017) 1606817. (2019) 869–877.
[46] X. Gao, X. Yang, M. Li, Q. Sun, J. Liang, J. Luo, J. Wang, W. Li, J. Liang, Y. Liu, [81] a) W.G. Lim, C. Jo, A. Cho, J. Hwang, S. Kim, J.W. Han, J. Lee, Adv. Mater. 31
S. Wang, Y. Hu, Q. Xiao, R. Li, T.K. Sham, X. Sun, Adv. Funct. Mater. 29 (2019) (2019) 1806547.
1806724. [82] S.S. Huang, M.T. Tung, C.D. Huynh, B.J. Hwang, P.M. Bieker, C.C. Fang, N.L. Wu,
[47] L. Luo, S.H. Chung, A. Manthiram, J. Mater. Chem. A 6 (2018) 7659–7667. A.C.S. Sustain, Chem. Eng. 7 (2019) 7851–7861.
[48] X. Huang, J. Tang, B. Luo, R. Knibbe, T. Lin, H. Hu, M. Rana, Y. Hu, X. Zhu, Q. Gu, [83] C.C. Chuang, Y.Y. Hsieh, W.C. Chang, H.Y. Tuan, Chem. Eng. J. 387 (2020),
D. Wang, L. Wang, Adv. Energy Mater. 9 (2019) 1901872. 123904.
[49] Z. Wang, J. Shen, J. Liu, X. Xu, Z. Liu, R. Hu, L. Yang, Y. Feng, J. Liu, Z. Shi, [84] Z. Du, X. Chen, W. Hu, C. Chuang, S. Xie, A. Hu, W. Yan, X. Kong, X. Wu, H. Ji, L.
L. Ouyang, Y. Yu, M. Zhu, Adv. Mater. 31 (2019) 1902228. J. Wan, J. Am. Chem. Soc. 141 (2019) 3977–3985.
[50] Y.T. Liu, D.D. Han, L. Wang, G.R. Li, S. Liu, X.P. Gao, Adv. Energy Mater. 9 (2019) [85] P. Han, S.H. Chung, A. Manthiram, Small 15 (2019) 1900690.
1803477. [86] C. Zhou, M. Hong, Y. Yang, C. Yang, N. Hu, L. Zhang, Z. Yang, Y. Zhang, J. Power
[51] Y. Chen, W. Zhang, D. Zhou, H. Tian, D. Su, C. Wang, D. Stockdale, F. Kang, B. Li, Sources 438 (2019), 227044.
G. Wang, ACS Nano 13 (2019) 4731–4741. [87] D. Moy, A. Manivannan, S.R. Narayanan, J. Electrochem. Soc. 162 (2014) A1–A7.
[52] Z. Li, Q. He, X. Xu, Y. Zhao, X. Liu, C. Zhou, D. Ai, L. Xia, L. Mai, Adv. Mater. 30 [88] W. Chen, T. Lei, W. Lv, Y. Hu, Y. Yan, Y. Jiao, W. He, Z. Li, C. Yan, J. Xiong, Adv.
(2018) 1804089. Mater. 30 (2018) 1804084.
[53] X. Li, B. Gao, X. Huang, Z. Guo, Q. Li, X. Zhang, P.K. Chu, K. Huo, ACS Appl. Mater.
Interf. 11 (2019) 2961–2969.

12

You might also like