Ultrathin Honeycomb-like Carbon as Sulfur Host ... - ACS Publications

Nov 26, 2018 - Here an ultrathin honeycomb-like porous carbon derived from loofah sponge ... Attributed to the integrated honeycomb structure, hierarc...
0 downloads 0 Views 5MB Size
Article Cite This: ACS Appl. Energy Mater. 2018, 1, 7076−7084

www.acsaem.org

Ultrathin Honeycomb-like Carbon as Sulfur Host Cathode for High Performance Lithium−Sulfur Batteries Taohong Wang,†,‡ Yuhua Yang,† Ling Fan,† Longlu Wang,† Ruifang Ma,† Qingfeng Zhang,† Jianguo Zhao,†,⊥ Junmin Ge,† Xianlu Lu,† Xinzhi Yu,† Hongguan Yang,*,† and Bingan Lu*,†,§,∥ †

School of Physics and Electronics, Hunan University, Changsha 410082, China Physics Department, College of Science, Central South University of Forestry and Technology, Changsha 410004, China § Fujian Strait Research Institute of Industrial Graphene Technologies, Jinjiang 362200, China ∥ State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan, Shanxi 030001, China ⊥ School of Physics and Electronic Information, Luoyang Normal University, Luoyang, Henan 471022, P. R. China

Downloaded via UNIV OF SOUTH DAKOTA on December 24, 2018 at 14:27:01 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: Lithium−sulfur (Li−S) batteries have attracted great interest because of its high theoretical capacity (1675 mA h g−1). However, the low electrical conductivity of sulfur, dissolution of polysulfides, and structural collapse of electrode limit its practical application. Here an ultrathin honeycomb-like porous carbon derived from loofah sponge and doped with nitrogen (PCLSN) is prepared as a stable host for sulfur nanoparticles. Attributed to the integrated honeycomb structure, hierarchical porosity, ultrathin honeycomb walls, and synergistic effects between physical and chemical adsorption of polysulfides, the developed PCLSN@S cathode achieves a high initial specific capacity of 1379 mA h g−1 at 0.1 C, outstanding cyclability with a small capacity decay rate of 0.044% per cycle over 970 cycles at 2 C, excellent rate performance with a high capacity of 664 mA h g−1 at 3 C, and high sulfur content of PCLSN@S up to 76.1%. Our approach provides a promising route to design other 3D porous structures for high performance Li−S batteries. KEYWORDS: lithium−sulfur batteries, ultrathin, honeycomb-like, porous carbon, loofah sponge



INTRODUCTION With global warming and air pollution caused by the consumption of fossil fuels, it is necessary to develop lowcost and environmentally friendly energy-storage systems.1−5 Lithium−sulfur (Li−S) battery is considered one of the most promising candidates for the demand owing to its relatively nonpoisonous materials, high natural abundance, high theoretical energy density (2600 W h kg−1), and the outstanding theoretical capacity of sulfur (1675 mA h g−1 or 2800 W h L−1).6−10 Unfortunately, there are still many challenges impeding the practical utilization of Li−S batteries.11−13 First, the poor electrical conductivity of both sulfur and its final discharge products lead to the low utilization of active material. Second, the dissolution of polysulfide intermediates (Li2Sx, x = 3−8) generated in the cathode side © 2018 American Chemical Society

can shuttle between the cathode and anode and react with Li metal chemically, resulting in active materials loss and dendrite issues of Li anode, causing fast capacity degradation and safety hazards.14−19 Moreover, a volumetric expansion during electrochemical lithiation and delithiation processes results in the structural collapse of the electrode.20−32 The solutions to these issues are straightforward: improve the cathode conductivity and stabilize the active material within the cathode. Carbonaceous materials, with a high conductivity, tunable pore structure, and high stability, play a particularly crucial role in the sulfur electrode.23,33 Various Received: September 6, 2018 Accepted: November 26, 2018 Published: November 26, 2018 7076

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials

Figure 1. Schematic illustration of the synthetic procedure of PCLSN@S.

Figure 2. SEM images of (a) PCLS, (b) PCLSN, and (c−e) PCLSN@S. TEM images of (f, g) LSPCN and (h) LSPCN@S. (i) HAADF-STEM image of PCLSN@S and the corresponding EDS elemental mappings of C, N, and S.

carbon scaffolds, including carbon black,34 activated carbons,35,36 carbon nanotubes,37−40 and graphene22,41−43 have been applied to obtain carbon/sulfur composite cathodes in recent years. Although progress resulted, the cycling stability and discharge/charge capacity are still not satisfactory. The

above-mentioned carbon materials tend to pack together, leading to inhomogeneous contacts and poor links between the active material and the conducting matrix, which limit the utilization of the active material.44−47 Moreover, carbon nanotubes and graphene heavily rely on natural graphite 7077

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials

Figure 3. (a) XRD patterns of PCLS, PCLSN, PCLSN@S, and sulfur. (b) Raman spectra of PCLS, PCLSN, and PCLSN@S. (c) Isotherms and (d) pore size distributions of PCLSN.



RESULTS AND DISCUSSION Synthesis of PCLSN@S Nanostructures. Preparation of PCLSN@S structures is schematically shown in Figure 1. Briefly, dry loofah sponge was first soaked in potassium hydroxide (KOH) aqueous solution few days, followed by freeze-drying and calcination in argon atmosphere at 800 °C. During the annealing process, KOH served as a chemical activating reagent to obtain porous carbon material,52,58−60 which was derived from loofah sponge, denoted as PCLS. The obtained PCLS structure was washed with 1 M dilute hydrochloric acid (HCl) solution and deionized water. After freeze-drying, the PCLS material annealed in NH3 atmosphere at 500 °C to form nitrogen doped PCLS product, referred as PCLSN. Finally, sulfur was impregnated into the PCLSN structure by melt-diffusion at 155 °C for 24 h and 200 °C for 2 h to yield PCLSN@S composite. Morphology and Structure of PCLSN@S. Figure 2a and Figure S1a illustrate that the KOH etched PCLS product possesses a macroporous structure; the size of the holes is about 0.5−3.0 μm. As a comparison, Figure S1b and Figure S1c (Supporting Information) show the framework of the carbon from loofah sponge before being handled by KOH (CLS). By comparing the morphologies between PCLS and CLS, we can conclude that the surface of CLS is nonporous, the surface of PCLS is porous, which indicates that KOH is typically used as a pore-forming agent to induce a high porosity on PCLS. Figure 2b and Figure S1d display typical SEM overview of the PCLSN powders, which are evidently made of honeycomblike integrated structure; the size of the holes is maintained at about 0.5−3.0 μm. Clearly, after nitrogen doping by calcination in NH3, cavities on the surface of PCLSN material are overlapped to form 3D interconnected honeycomb

mines which are not cost-effective, restricting their wide application.48 Besides, for the nonpolar carbon framework, polar sulfur/lithium polysulfides may detach from it during discharge and charge processes due to the low binding energy between them, which results in severe shuttle effect and capacity decay.49 Heteroatom-doped carbon that has a polar nature is commonly used to trap the soluble lithium polysulfides from diffusing out the cathode through strong chemical binding between polysulfides molecules and carbon surface.50−52 Herein, we successfully developed an ultrathin honeycomblike hierarchical porous carbon material derived from loofah sponge and doped with nitrogen (PCLSN), as highly efficient sulfur host for Li−S batteries. Benefiting from the unique structure, PCLSN exhibits several advantages. First, the honeycomb structure requires minimum materials and has the highest density and maximum space; these properties are expected to incorporate more sulfur, increasing the loading and utilization of sulfur in Li−S batteries.53−56 Second, the interconnected network structure with micro/meso/macro porosity can facilitate fast electron/ion transport as well as effective physical trapping of polysulfides during the sulfur redox reactions. Finally, the nitrogen doping can further enhance chemical adsorption of polysulfides, resulting in synergistic effects between physical and chemical adsorption of polysulfides.33,57 As a result, the PCLSN based cathode provides the Li−S cell with significant enhancements. The outstanding cell performance is evidenced by a high discharge capacity of 1379 mA h g−1 at 0.1 C, a low capacity fade rate of 0.044% per cycle for 970 cycles at 2 C, and excellent rate performance from 0.1 to 3 C. 7078

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials

Figure 4. (a) C 1s and (b) N 1s XPS spectra for PCLSN. (c) S 2p XPS spectrum for PCLSN@S. (d) TGA curve of PCLSN@S in Ar atmosphere.

structure, which facilitates an easy access to Li+ ingress/egress. Besides, micro/mesopores of PCLS and PCLSN can be seen from Figure S1e and S1f (Supporting Information). The micro/meso/macro pores are ascribed to the etching of carbon framework by redox reactions and gasification of carbon during KOH activation process and the removal of metallic K and other K compounds produced in KOH activation process by washing with HCl dilute solution and deionized water.61 The SEM images of the PCLSN@S reveal that sulfur is uniformly dispersed in cavities of PCLSN, and there still is a considerable amount of void holes on the surface of PCLSN@S (Figure 2c− e) to accommodate the large volumetric expansion, thus keeping the structural integrity to reduce polysulfides dissolution. The honeycomb-like PCLSN structure could be further evidenced by low-magnification TEM image in Figure 2f, which was consistent with SEM images. The high-magnification TEM image of PCLSN in Figure 2g reveals the walls of honeycombs can be thin to 6 nm; the ultrathin walls of honeycombs allow easier transit for Li+ ion than thick physical barriers,62 which leads to excellent conductivity. Besides, PCLSN has a degree of graphitization, which further benefits the electronic conductivity of carbon.63 The TEM observation of PCLSN@S (Figure 2h) reals that the 3D network structure was maintained intact after sulfur infused into PCLSN. Figure 2i presents the high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) image and the corresponding energy-dispersive X-ray spectroscopy (EDS) elemental mapping; the mapping confirmed the presence of carbon, nitrogen and sulfur. Besides, the nitrogen and sulfur were uniformly distributed along the carbon. The structural characteristics of PCLSN@S nanocomposite are detected by X-ray diffraction (XRD), as shown in Figure 3a. The XRD pattern of the PCLSN@S shows two broad

diffraction peaks, which are attributed to the (002) and (100) planes of carbon (JCPDS No. 75-1621) originating from PCLS and PCLSN carbon materials.64−66 Strong characteristic peaks of orthorhombic sulfur (JCPDS No. 08-0247) in PCLSN@S indicate the successful incorporation of sulfur into PCLSN carbon.67 Figure 3b shows the Raman spectral analysis of PCLS, PCLSN, and PCLSN@S. Two typical peaks were found at about 1335 and 1589 cm−1, which are respectively assigned to the disorder-induced D band and G band arising from the symmetric stretching mode of C−C bonds from the ordered structure of graphitic crystallites of carbon.68−70 The intensity ratio of D band and G band (ID/IG) of the PCLS, PCLSN, and PCLSN@S are 1.01, 1.03, and 1.08, respectively, manifesting the carbon matrix is partially graphitized, consistent with the TEM images. The characteristic peaks of sulfur were not observed in Figure 3b, which may be ascribed to the full diffusion of sulfur into pores of PCLSN. Figure 3c shows the nitrogen adsorption−desorption isotherm curves of the PCLSN; the Brunauer−Emmett−Teller (BET) specific surface and total pore volume of the PCLSN are 1033.0 m2 g−1 and 0.68 cm3 g−1, respectively. The pore size distribution of the PCLSN shown in Figure 3d indicates a micro/mesoporous structure in the range between 1.2 and 5.0 nm with an average diameter of 2.64 nm (Table S1). The high porosity and large surface area in the PCLSN composites afford abundant space for loading of active materials and provide open channels for electrolyte penetration. The nitrogen adsorption−desorption isotherm curves of the PCLS was shown in Figure S2. X-ray photoelectron spectroscopy (XPS) was carried out to confirm the chemical composition of PCLSN (Figure 4a−c). In the high-resolution C 1s spectrum (Figure 4a), there are three peaks at 284.8, 285.7 and 287.6 eV, which were assigned 7079

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials

Figure 5. (a) CV curves of the honeycomb-like PCLSN@S electrode at a scan rate of 0.1 mV s−1. (b) Voltage-capacity profiles and (c) rate performance of PCLSN@S electrode at various current densities from 0.1 to 3 C. (d) Cycling performance of PCLSN@S electrode at 2 C. The Coulombic efficiency is nearly 100%.

During the first cathodic scan, the two reduction peaks centered at 2.27 and 1.99 V are attributed to the reduction of S8 to long-chain polysulfides (Li2Sx, 4 ≤ x ≤ 8) and the further reduction of the long-chain polysulfides to insoluble Li2S2/ Li2S, respectively.75 In the subsequent anodic scan, a main oxidation peak at 2.39 V corresponds to the reversible processes from Li2S to Li2Sx and ultimately to S8. In the third cycle, the two reduction peaks shift to higher potentials at 2.28 and 2.00 V, respectively, and the anodic peak shifts to lower potential at 2.35 V, suggesting a small electrochemical polarization.76 No significant changes of the curve between the third and fifth cycle indicate that electrochemical performance tends to be stable. Figure 5b reports the discharge/charge profiles at different current rates from 0.1 to 3.0 C (1 C = 1675 mA g−1). A high initial discharge capacity of 1379 mA h g−1 was achieved at 0.1 C, indicating high sulfur utilization. The discharge curves exhibit two typical plateaus around 2.3 and 2.1 V, corresponding to the reduction of sulfur to long chain polysulfides and the formation of insoluble Li2S2/Li2S, being well consistent with the CVs. The rate capability of PCLSN@S electrode was assessed under current densities ranging from 0.1 to 3.0 C, as shown in Figure 5c. When cycled at 0.1, 0.3, 0.5, 1.0, 2.0, and 3.0 C, the PCLSN@S cathode can deliver impressive discharge capacities of 1379, 1104, 1006, 925, 805, and 664 mA h g−1, respectively. As the current density switches back to 0.1 C, the discharge

to the C−C, C−N and C−O bonding, respectively,71 suggesting that the coexistence of carbon, nitrogen and a small amount of oxygen in PCLSN. Figure 4b shows the fitted N 1s spectrum, which can be divided into three peaks at 398.7, 399.9 and 401.0 eV, corresponding to the pyridine N, pyrrole N and graphitic N, respectively;71,72 the pyridinic- and pyrrolic-type N are dominant as shown in Figure 4b. The content of nitrogen in PCLSN is 2.2 wt %; these various types of doped nitrogen atoms increase the polarity of carbon surface, which utilizes chemical adsorption to mitigate the shuttling effect.73 The S 2p XPS spectrum of PCLSN@S is resolved into S 2p3/2 and S 2p1/2, two peaks at 163.8 and 165.0 eV in Figure 4c, respectively. The difference in binding energy between S 2p3/2 and S 2p1/2 is 1.2 eV, and the intensity ratio is about 2:1.74 The C 1s and N 1s XPS spectra of PCLSN@S were shown in Figure S3. The weight ratio of sulfur in PCLSN@S was determined to be 76.1% according to the thermogravimetric analysis (TGA) curves, as shown in Figure 4d. The high loading of sulfur benefits from the high porosity and nitrogen atoms adoption for PCLSN. Electrochemical Performance. The electrochemical performance of PCLSN@S composite is evaluated as a cathode material for Li−S batteries; the sulfur loading of electrode is about 1 mg cm−2, and all the capacities were calculated based on the mass of sulfur in electrode. Figure 5a displays the cyclic voltammetry (CV) curves of the cell at a scan of 0.02 mV s−1 in the potential range of 1.6−2.8 V. 7080

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials capacity is restored to 1097 mA h g−1, indicating excellent electrochemical reversibility. As displayed in Figure 5d, the PCLSN@S nanoarchitecture exhibited stable cycling performance over 970 cycles at 2 C. A reversible specific capacity of 406 mA h g−1 was retained with capacity retentions of 60.4% and capacity decay of 0.044% per cycle for 970 cycles. The average Coulombic efficiency was 99.6% over 970 cycles, nearly 100%, which shows little shuttle effect and high cycling stability.77 To show the excellent properties of PCLSN@S cathode, the electrochemical performance of CLS@S and PCLS@S cathodes is also investigated (Figures S4−S7). Furthermore, Table S2 shows that the electrochemical performance of PCLSN@S cathode is better than most of the reported lithium sulfur batteries. This outstanding performance for Li−S batteries based on the PCLSN composite should be attributed to the following factors. First, the 3D interwoven honeycomb-like carbon framework with abundant micro/meso/macro pores not only provides the fast ion and electron transfer channels but also suppresses the shuttling effect during the discharge/charge cycles by physical confinement. Second, nitrogen functional groups further increase the adsorption toward the Li polysulfides by chemical bonding for prolonged cycling life. The synergy effects between the physical and chemical adsorption improve the sulfur utilization. Third, the porous structure can present sufficient empty space against the large volumetric expansion without causing the structure crack and fracture. Furthermore, the ultrathin and graphitization characteristics of the carbon improve the electronic conductivity, thus obtaining high rate capability.75,77,78

mixing PCLSN and sulfur (1:4 by mass ratio) by strongly grinding for 20 min and then coheating to 110 °C with 30 min, continued heating to 155 °C with a heating rate of 0.5 °C min−1 for 24 h, and last, continued heating to 200 °C with 30 min for 2 h under argon protection in tube furnace. Characterization. The samples were characterized by SEM (Hitachi, S-4800, 5 kV), TEM (Titan G2 60-300), XRD (Philips, X’ pertpro), Raman spectroscopy (Renishaw 2000 system), XPS (ESCALAB 250Xi), BET (Quabrasorb SI-3MP), and TGA (Netzsch STA 449). Electrochemical Measurements. The electrode slurry was prepared by mixing PCLSN@S composites with acetylene black and polyvinylidene fluoride (PVDF) at a weight ratio of 7:2:2 in Nmethyl-2-pyrrolidone (NMP). The electrode with an average loading of sulfur of about 1 mg cm−2 and a thickness of 18 um (Figure S8) was prepared by coating the slurry onto an aluminum foil and dried at 60 °C for 24 h in vacuum. 2032-type coin cells were applied to assemble test cells. The electrolyte was 1 M bis(trifluoromethane)sulfonamide lithium salt (LiTFSi) in a mixed solvent of 1,3-dioxolane (DOL) and 1,2-dimethoxyethane (DME) with a volume ratio of 1:1, including 1 wt % LiNO3 as an electrolyte additive, the amount of electrolyte is excessive. Lithium metal foil with a thickness of 0.58 mm was used as the counter electrode, and a microporous polypropylene acted as the separator. 2032-type coin cells were assembled in a glovebox filled with argon (H2O < 0.5 ppm, O2 < 0.5 ppm). For comparison, CLS@S, PCLS@S cathode with sulfur loading of about 1 mg cm−2, and PCLSN@S cathode with sulfur loading of about 4 mg cm−2 (Figure S9) were prepared using the same method.



S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsaem.8b01498. SEM images of CLS, PCLS, and PCLSN, TEM images of PCLS and PCLSN, BET curves of PCLS, XPS curves of PCLSN@S, voltage-capacity profiles of CLS@S and PCLS@S, cycling performance of CLS@S and PCLS@S, SEM pictures of the PCLSN@S cathode electrode, cycling performance of PCLSN@S electrode with a sulfur loading of 4 mg cm−2, table for N2 adsorption− desorption isothermal analysis of the PCLS and PCLSN, table for electrochemical performances of carbon materials-based cathodes for Li−S batteries (PDF)



CONCLUSIONS We have rationally designed an integrated ultrathin honeycomb-like PCLSN carbon with hierarchical porous structure as a highly effective sulfur host for Li−S batteries. The strategy utilizes dry loofah sponge as the carbon precursor, KOH as the pore-forming agent, and NH3 as the nitrogen dopant. With the help of this 3D cross-linked honeycomb structure, the Li−S batteries exhibited high sulfur utilization up to 76.1%, high specific capacities of 1379 mA h g−1 at 0.1 C, good rate capabilities from 0.1 to 3 C, especially excellent cycling stabilities with a capacity decay of only 0.044% per cycle over 970 cycles at 2 C current. This work provides a promising route to design other 3D porous structures for highperformance Li−S batteries.



ASSOCIATED CONTENT



AUTHOR INFORMATION

Corresponding Authors

*H.Y.: e-mail, [email protected]. *B.L.: e-mail, [email protected]. ORCID

EXPERIMENTAL SECTION

Ling Fan: 0000-0002-4813-3895 Bingan Lu: 0000-0002-0075-5898

Materials. Dry loofah sponge was bought from Taobao. All other reagents were of analytical grade and used without further purification. Synthesis of CLS, PCLS, PCLSN, and PCLSN@S Nanostructures. 10 g dry loofah sponge was first cut into pieces and soaked in 200 mL of 3 M potassium hydroxide (KOH) aqueous solution for 10 days at room temperature. Subsequently, the loofah sponge was collected and freeze-dried for 2 days. Then, the KOH treated loofah sponge was heated at 800 °C for 2 h with a heating rate of 5 °C min−1 under argon protection in a tube furnace. The resultant product was washed with 1 M dilute hydrochloric acid (HCl) solution and deionized water three times. Finally, the PCLS sample was obtained. For comparison, CLS was also prepared using dry loofah sponge without KOH treatment through the same process. After freezedrying, the obtained PCLS material was annealing in NH3 atmosphere at 500 °C for 2 h with a heating rate of 5 °C min−1. The product was referred to as PCLSN. Finally, the PCLSN@S composite was made by

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by National Natural Science Foundation of China (Grants 51672078 and 21473052), Hunan University State Key Laboratory of Advanced Design and Manufacturing for Vehicle Body Independent Research Project (Grant 71675004), Hunan Youth Talents (Grant 2016RS3025), Foundation of State Key Laboratory of Coal Conversion (Grant J17-18-903), National Natural Science Foundation of China (Grants 51672078, 21473052 and 61474041), and the Program for Innovation 7081

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials

Broadband MoS(2) Photodetector Driven by Ferroelectrics. Adv. Mater. 2015, 27, 6575−81. (19) Yasaei, P.; Kumar, B.; Foroozan, T.; Wang, C.; Asadi, M.; Tuschel, D.; Indacochea, J. E.; Klie, R. F.; Salehi-Khojin, A. HighQuality Black Phosphorus Atomic Layers by Liquid-Phase Exfoliation. Adv. Mater. 2015, 27, 1887−1892. (20) Tao, T.; Lu, S.; Fan, Y.; Lei, W.; Huang, S.; Chen, Y. Anode Improvement in Rechargeable Lithium−Sulfur Batteries. Adv. Mater. 2017, 29, 1700542. (21) Zhou, W.; Yu, Y.; Chen, H.; DiSalvo, F. J.; Abruña, H. c. D. Yolk-Shell Structure of Polyaniline-Coated Sulfur for Lithium-Sulfur Batteries. J. Am. Chem. Soc. 2013, 135, 16736−16743. (22) Huang, J. Q.; Liu, X. F.; Zhang, Q.; Chen, C. M.; Zhao, M. Q.; Zhang, S. M.; Zhu, W.; Qian, W. Z.; Wei, F. Entrapment of sulfur in hierarchical porous graphene for lithium−sulfur batteries with high rate performance from −40 to 60 °C. Nano Energy 2013, 2, 314−321. (23) Fang, R.; Li, G.; Zhao, S.; Yin, L.; Du, K.; Hou, P.; Wang, S.; Cheng, H. M.; Liu, C.; Li, F. Single-wall carbon nanotube network enabled ultrahigh sulfur-content electrodes for high-performance lithium-sulfur batteries. Nano Energy 2017, 42, 205−214. (24) Li, Z.; Guan, B. Y.; Zhang, J.; Lou, X. W. D. A Compact Nanoconfined Sulfur Cathode for High-Performance Lithium-Sulfur Batteries. Joule 2017, 1, 576−587. (25) Peled, E.; Sternberg, Y.; Gorenshtein, A.; Lavi, Y. LithiumSulfur Battery: Evaluation of Dioxolane-Based Electrolytes. J. Electrochem. Soc. 1989, 136, 1621−1625. (26) Lee, D. H.; Ahn, J. H.; Park, M.-S.; Eftekhari, A.; Kim, D.-W. Metal-organic framework/carbon nanotube-coated polyethylene separator for improving the cycling performance of lithium-sulfur cells. Electrochim. Acta 2018, 283, 1291−1299. (27) Fan, L.; Chen, S.; Zhu, J.; Ma, R.; Li, S.; Podila, R.; Rao, A. M.; Yang, G.; Wang, C.; Liu, Q.; Xu, Z.; Yuan, L.; Huang, Y.; Lu, B. Simultaneous Suppression of the Dendrite Formation and Shuttle Effect in a Lithium-Sulfur Battery by Bilateral Solid Electrolyte Interface. Adv. Sci. 2018, 5, 1700934. (28) Dai, C.; Hu, L.; Wang, M.-Q.; Chen, Y.; Han, J.; Jiang, J.; Zhang, Y.; Shen, B.; Niu, Y.; Bao, S.-J.; Xu, M. Uniform α-Ni(OH) 2 hollow spheres constructed from ultrathin nanosheets as efficient polysulfide mediator for long-term lithium-sulfur batteries. Energy Storage Mater. 2017, 8, 202−208. (29) Hu, L.; Dai, C.; Lim, J. M.; Chen, Y.; Lian, X.; Wang, M.; Li, Y.; Xiao, P.; Henkelman, G.; Xu, M. A highly efficient double-hierarchical sulfur host for advanced lithium-sulfur batteries. Chem. Sci. 2018, 9, 666−675. (30) Cui, Y.; Zhang, Q.; Wu, J.; Liang, X.; Baker, A. P.; Qu, D.; Zhang, H.; Zhang, H.; Zhang, X. Developing porous carbon with dihydrogen phosphate groups as sulfur host for high performance lithium sulfur batteries. J. Power Sources 2018, 378, 40−47. (31) Gu, X.; Tong, C. J.; Rehman, S.; Liu, L. M.; Hou, Y.; Zhang, S. Multifunctional Nitrogen-Doped Loofah Sponge Carbon Blocking Layer for High-Performance Rechargeable Lithium Batteries. ACS Appl. Mater. Interfaces 2016, 8, 15991−6001. (32) Yang, J.; Chen, F.; Li, C.; Bai, T.; Long, B.; Zhou, X. A freestanding sulfur-doped microporous carbon interlayer derived from luffa sponge for high performance lithium−sulfur batteries. J. Mater. Chem. A 2016, 4, 14324−14333. (33) Chung, S.-H.; Manthiram, A. Carbonized Eggshell Membrane as a Natural Polysulfide Reservoir for Highly Reversible Li-S Batteries. Adv. Mater. 2014, 26, 1360−1365. (34) Shim, J.; Striebel, K. A.; Cairns, E. J. The Lithium/Sulfur Rechargeable Cell Effects of Electrode Composition and Solvent on Cell Performance. J. Electrochem. Soc. 2002, 149, A1321−A1325. (35) Wang, J. L.; Yang, J.; Xie, J. Y.; Xu, N. X.; Li, Y. Sulfur−carbon nano-composite as cathode for rechargeable lithium battery based on gel electrolyte. Electrochem. Commun. 2002, 4, 499−502. (36) Wang, J.; Liu, L.; Ling, Z.; Yang, J.; Wan, C.; Jiang, C. Polymer lithium cells with sulfur composites as cathode materials. Electrochim. Acta 2003, 48, 1861−1867.

Talents (in Science and Technology) in University of Henan Province (Grant 16HASTIT044).



REFERENCES

(1) Seh, Z. W.; Sun, Y.; Zhang, Q.; Cui, Y. Designing high-energy lithium−sulfur batteries. Chem. Soc. Rev. 2016, 45, 5605−5634. (2) Zhang, Q.; Wang, L.; Wang, J.; Yu, X.; Ge, J.; Zhang, H.; Lu, B. Semimetallic vanadium molybdenum sulfide for high-performance battery electrodes. J. Mater. Chem. A 2018, 6, 9411−9419. (3) Zhang, S.; Ueno, K.; Dokko, K.; Watanabe, M. Recent Advances in Electrolytes for Lithium-Sulfur Batteries. Adv. Energy Mater. 2015, 5, 1500117. (4) Xu, X.; Liu, J.; Liu, J.; Ouyang, L.; Hu, R.; Wang, H.; Yang, L.; Zhu, M. A General Metal-Organic Framework (MOF)-Derived Selenidation Strategy for In Situ Carbon-Encapsulated Metal Selenides as High-Rate Anodes for Na-Ion Batteries. Adv. Funct. Mater. 2018, 28, 1707573. (5) Xu, X.; Liu, J.; Liu, Z.; Shen, J.; Hu, R.; Liu, J.; Ouyang, L.; Zhang, L.; Zhu, M. Robust Pitaya-Structured Pyrite as High Energy Density Cathode for High-Rate Lithium Batteries. ACS Nano 2017, 11, 9033−9040. (6) Fang, R.; Zhao, S.; Sun, Z.; Wang, D. W.; Cheng, H. M.; Li, F. More Reliable Lithium-Sulfur Batteries: Status, Solutions and Prospects. Adv. Mater. 2017, 29, 1606823. (7) Chen, W.; Lei, T.; Qian, T.; Lv, W.; He, W.; Wu, C.; Liu, X.; Liu, J.; Chen, B.; Yan, C.; Xiong, J. A New Hydrophilic Binder Enabling Strongly Anchoring Polysulfides for High-Performance Sulfur Electrodes in Lithium-Sulfur Battery. Adv. Energy Mater. 2018, 8, 1702889. (8) Yamin, H.; Gorenshtein, A.; Penciner, J.; Sternberg, Y.; Peled, E. Lithium Sulfur Battery. Oxidation/Reduction Mechanisms of Polysulfides in THF Solutions. J. Electrochem. Soc. 1988, 135, 1045−1048. (9) Hu, L.; Dai, C.; Liu, H.; Li, Y.; Shen, B.; Chen, Y.; Bao, S.-J.; Xu, M. Double-Shelled NiO-NiCo2O4 Heterostructure@Carbon Hollow Nanocages as an Efficient Sulfur Host for Advanced Lithium-Sulfur Batteries. Adv. Energy Mater. 2018, 8, 1800709. (10) Dai, C.; Hu, L.; Li, X.; Xu, Q.; Wang, R.; Liu, H.; Chen, H.; Bao, S.-J.; Chen, Y.; Henkelman, G.; Li, C. M.; Xu, M. Chinese knotlike electrode design for advanced Li-S batteries. Nano Energy 2018, 53, 354−361. (11) Zhou, X.; Chen, F.; Yang, J. Core@shell sulfur@polypyrrole nanoparticles sandwiched in graphene sheets as cathode for lithium− sulfur batteries. J. Energy Chem. 2015, 24, 448−455. (12) Ding, H. B.; Zhang, Q. F.; Liu, Z. M.; Wang, J.; Ma, R. F.; Fan, L.; Wang, T.; Zhao, J. G.; Ge, J. M.; Lu, X. L.; Yu, X. Z.; Lu, B. G. TiO2 quantum dots decorated multi-walled carbon nanotubes as the multifunctional separator for highly stable lithium sulfur batteries. Electrochim. Acta 2018, 284, 314−320. (13) Dai, C.; Lim, J.-M.; Wang, M.; Hu, L.; Chen, Y.; Chen, Z.; Chen, H.; Bao, S.-J.; Shen, B.; Li, Y.; Henkelman, G.; Xu, M. Honeycomb-Like Spherical Cathode Host Constructed from Hollow Metallic and Polar Co9S8 Tubules for Advanced Lithium-Sulfur Batteries. Adv. Funct. Mater. 2018, 28, 1704443. (14) Cheng, X.-B.; Huang, J.-Q.; Zhang, Q. ReviewLi Metal Anode in Working Lithium-Sulfur Batteries. J. Electrochem. Soc. 2018, 165, A6058−A6072. (15) Li, W.; Yao, H.; Yan, K.; Zheng, G.; Liang, Z.; Chiang, Y. M.; Cui, Y. The synergetic effect of lithium polysulfide and lithium nitrate to prevent lithium dendrite growth. Nat. Commun. 2015, 6, 7436. (16) Yan, C.; Cheng, X.-B.; Zhao, C.-Z.; Huang, J.-Q.; Yang, S.-T.; Zhang, Q. Lithium metal protection through in-situ formed solid electrolyte interphase in lithium-sulfur batteries: The role of polysulfides on lithium anode. J. Power Sources 2016, 327, 212−220. (17) Liu, B.; Zhang, J.-G.; Xu, W. Advancing Lithium Metal Batteries. Joule 2018, 2, 833−845. (18) Wang, X.; Wang, P.; Wang, J.; Hu, W.; Zhou, X.; Guo, N.; Huang, H.; Sun, S.; Shen, H.; Lin, T.; Tang, M.; Liao, L.; Jiang, A.; Sun, J.; Meng, X.; Chen, X.; Lu, W.; Chu, J. Ultrasensitive and 7082

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials (37) Han, S.-C.; Song, M.-S.; Lee, H.; Kim, H.-S.; Ahn, H.-J.; Lee, J.Y. Effect of Multiwalled Carbon Nanotubes on Electrochemical Properties of Lithium/Sulfur Rechargeable Batteries. J. Electrochem. Soc. 2003, 150, A889−A893. (38) Cheng, X.-B.; Huang, J.-Q.; Zhang, Q.; Peng, H.-J.; Zhao, M.Q.; Wei, F. Aligned carbon nanotube/sulfur composite cathodes with high sulfur content for lithium−sulfur batteries. Nano Energy 2014, 4, 65−72. (39) Hagen, M.; Dörfler, S.; Althues, H.; Tübke, J.; Hoffmann, M. J.; Kaskel, S.; Pinkwart, K. Lithium−sulphur batteries − binder free carbon nanotubes electrode examined with various electrolytes. J. Power Sources 2012, 213, 239−248. (40) Sun, L.; Li, M.; Jiang, Y.; Kong, W.; Jiang, K.; Wang, J.; Fan, S. Sulfur nanocrystals confined in carbon nanotube network as a binderfree electrode for high-performance lithium sulfur batteries. Nano Lett. 2014, 14, 4044. (41) Zhou, G.; Li, L.; Ma, C.; Wang, S.; Shi, Y.; Koratkar, N.; Ren, W.; Li, F.; Cheng, H.-M. A graphene foam electrode with high sulfur loading for flexible and high energy Li-S batteries. Nano Energy 2015, 11, 356−365. (42) Chen, R.; Zhao, T.; Lu, J.; Wu, F.; Li, L.; Chen, J.; Tan, G.; Ye, Y.; Amine, K. Graphene-based three-dimensional hierarchical sandwich-type architecture for high-performance Li/S batteries. Nano Lett. 2013, 13, 4642. (43) Zhao, M. Q.; Zhang, Q.; Huang, J. Q.; Tian, G. L.; Nie, J. Q.; Peng, H. J.; Wei, F. Unstacked double-layer templated graphene for high-rate lithium-sulphur batteries. Nat. Commun. 2014, 5, 3410. (44) Schuster, J.; He, G.; Mandlmeier, B.; Yim, T.; Lee, K. T.; Bein, T.; Nazar, L. F. Spherical ordered mesoporous carbon nanoparticles with high porosity for lithium-sulfur batteries. Angew. Chem. 2012, 124, 3651−3655. (45) Tang, C.; Zhang, Q.; Zhao, M. Q.; Huang, J. Q.; Cheng, X. B.; Tian, G. L.; Peng, H. J.; Wei, F. Nitrogen-doped aligned carbon nanotube/graphene sandwiches: facile catalytic growth on bifunctional natural catalysts and their applications as scaffolds for high-rate lithium-sulfur batteries. Adv. Mater. 2014, 26, 6100−6105. (46) Peng, H.-J.; Huang, J.-Q.; Zhao, M.-Q.; Zhang, Q.; Cheng, X.B.; Liu, X.-Y.; Qian, W.-Z.; Wei, F. Nanoarchitectured Graphene/ CNT@Porous Carbon with Extraordinary Electrical Conductivity and Interconnected Micro/Mesopores for Lithium-Sulfur Batteries. Adv. Funct. Mater. 2014, 24, 2772−2781. (47) Manthiram, A.; Chung, S. H.; Zu, C. Lithium-sulfur batteries: progress and prospects. Adv. Mater. 2015, 27, 1980−2006. (48) Wang, C.; Lan, M.; Zhang, Y.; Bian, H.; Yuen, M.-F.; Ostrikov, K.; Jiang, J.; Zhang, W.; Li, Y. Y.; Lu, J. Fe1−xS/C nanocomposites from sugarcane waste-derived microporous carbon for high-performance lithium ion batteries. Green Chem. 2016, 18, 3029−3039. (49) Wu, F.; Ye, Y.; Chen, R.; Qian, J.; Zhao, T.; Li, L.; Li, W. Systematic Effect for an Ultralong Cycle Lithium-Sulfur Battery. Nano Lett. 2015, 15, 7431. (50) Bao, W.; Liu, L.; Wang, C.; Choi, S.; Wang, D.; Wang, G. Facile Synthesis of Crumpled Nitrogen-Doped MXene Nanosheets as a New Sulfur Host for Lithium-Sulfur Batteries. Adv. Energy Mater. 2018, 8, 1702485. (51) Zhang, S. S. Heteroatom-doped carbons: synthesis, chemistry and application in lithium/sulphur battery. Inorg. Chem. Front. 2015, 2, 1059−1069. (52) Ren, G.; Li, S.; Fan, Z.-X.; Warzywoda, J.; Fan, Z. Soybeanderived hierarchical porous carbon with large sulfur loading and sulfur content for high-performance lithium−sulfur batteries. J. Mater. Chem. A 2016, 4, 16507−16515. (53) Li, Y.; Fan, J.; Zhang, J.; Yang, J.; Yuan, R.; Chang, J.; Zheng, M.; Dong, Q. A Honeycomb-like Co@N-C Composite for Ultrahigh Sulfur Loading Li-S Batteries. ACS Nano 2017, 11, 11417−11424. (54) Xu, B.; Arias, F.; Whitesides, G. M. Making Honeycomb Microcomposites by Soft Lithography. Adv. Mater. 1999, 11, 492− 495.

(55) Ma, J.; Hui, Y. S.; Zhang, M.; Yu, Y.; Wen, W.; Qin, J. Facile synthesis of biomimetic honeycomb material with biological functionality. Small 2013, 9, 497−503. (56) Yu, X.; Lu, B.; Xu, Z. Super long-life supercapacitors based on the construction of nanohoneycomb-like strongly coupled CoMoO(4)-3D graphene hybrid electrodes. Adv. Mater. 2014, 26, 1044−1051. (57) Song, J.; Xu, T.; Gordin, M. L.; Zhu, P.; Lv, D.; Jiang, Y.-B.; Chen, Y.; Duan, Y.; Wang, D. Nitrogen-Doped Mesoporous Carbon Promoted Chemical Adsorption of Sulfur and Fabrication of HighAreal-Capacity Sulfur Cathode with Exceptional Cycling Stability for Lithium-Sulfur Batteries. Adv. Funct. Mater. 2014, 24, 1243−1250. (58) You, Y.; Zeng, W.; Yin, Y.-X.; Zhang, J.; Yang, C.-P.; Zhu, Y.; Guo, Y.-G. Hierarchically micro/mesoporous activated graphene with a large surface area for high sulfur loading in Li−S batteries. J. Mater. Chem. A 2015, 3, 4799−4802. (59) Gu, X.; Lai, C.; Liu, F.; Yang, W.; Hou, Y.; Zhang, S. A conductive interwoven bamboo carbon fiber membrane for Li−S batteries. J. Mater. Chem. A 2015, 3, 9502−9509. (60) Zhong, Y.; Xia, X.; Deng, S.; Zhan, J.; Fang, R.; Xia, Y.; Wang, X.; Zhang, Q.; Tu, J. Popcorn Inspired Porous Macrocellular Carbon: Rapid Puffing Fabrication from Rice and Its Applications in LithiumSulfur Batteries. Adv. Energy Mater. 2018, 8, 1701110. (61) Wang, J.; Kaskel, S. KOH activation of carbon-based materials for energy storage. J. Mater. Chem. 2012, 22, 23710. (62) Jiang, J.; Zhu, J.; Ai, W.; Wang, X.; Wang, Y.; Zou, C.; Huang, W.; Yu, T. Encapsulation of sulfur with thin-layered nickel-based hydroxides for long-cyclic lithium-sulfur cells. Nat. Commun. 2015, 6, 8622. (63) Xu, G. L.; Xu, Y. F.; Fang, J. C.; Peng, X. X.; Fu, F.; Huang, L.; Li, J. T.; Sun, S. G. Porous graphitic carbon loading ultra high sulfur as high-performance cathode of rechargeable lithium-sulfur batteries. ACS Appl. Mater. Interfaces 2013, 5, 10782−10793. (64) Fan, L.; Lu, B. Reactive Oxygen-Doped 3D Interdigital Carbonaceous Materials for Li and Na Ion Batteries. Small 2016, 12, 2783−2791. (65) Fan, L.; Ma, R.; Yang, Y.; Chen, S.; Lu, B. Covalent sulfur for advanced room temperature sodium-sulfur batteries. Nano Energy 2016, 28, 304−310. (66) Fan, L.; Liu, Q.; Chen, S.; Xu, Z.; Lu, B. Soft Carbon as Anode for High-Performance Sodium-Based Dual Ion Full Battery. Adv. Energy Mater. 2017, 7, 1602778. (67) Kaiser, M. R.; Ma, Z.; Wang, X.; Han, F.; Gao, T.; Fan, X.; Wang, J.-Z.; Liu, H. K.; Dou, S.; Wang, C. Reverse Microemulsion Synthesis of Sulfur/Graphene Composite for Lithium/Sulfur Batteries. ACS Nano 2017, 11, 9048−9056. (68) Jiang, J.; Zhu, J.; Ai, W.; Fan, Z.; Shen, X.; Zou, C.; Liu, J.; Zhang, H.; Yu, T. Evolution of disposable bamboo chopsticks into uniform carbon fibers: a smart strategy to fabricate sustainable anodes for Li-ion batteries. Energy Environ. Sci. 2014, 7, 2670−2679. (69) Xiang, M.; Wu, H.; Liu, H.; Huang, J.; Zheng, Y.; Yang, L.; Jing, P.; Zhang, Y.; Dou, S.; Liu, H. A Flexible 3D Multifunctional MgODecorated Carbon Foam@CNTs Hybrid as Self-Supported Cathode for High-Performance Lithium-Sulfur Batteries. Adv. Funct. Mater. 2017, 27, 1702573. (70) Guo, J.; Xu, Y.; Wang, C. Sulfur-impregnated disordered carbon nanotubes cathode for lithium-sulfur batteries. Nano Lett. 2011, 11, 4288−4294. (71) Zhou, G.; Paek, E.; Hwang, G. S.; Manthiram, A. Long-life Li/ polysulphide batteries with high sulphur loading enabled by lightweight three-dimensional nitrogen/sulphur-codoped graphene sponge. Nat. Commun. 2015, 6, 7760. (72) Sheng, Z.-H.; Shao, L.; Chen, J.-J.; Bao, W.-J.; Wang, F.-B.; Xia, X.-H. Catalyst-Free Synthesis of Nitrogen-Doped Graphene via Thermal Annealing Graphite Oxide with Melamine and Its Excellent Electrocatalysis. ACS Nano 2011, 5, 4350−4358. (73) Zheng, C.; Niu, S.; Lv, W.; Zhou, G.; Li, J.; Fan, S.; Deng, Y.; Pan, Z.; Li, B.; Kang, F.; Yang, Q.-H. Propelling polysulfides 7083

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084

Article

ACS Applied Energy Materials transformation for high-rate and long-life lithium−sulfur batteries. Nano Energy 2017, 33, 306−312. (74) Wang, Z.; Dong, Y.; Li, H.; Zhao, Z.; Wu, H. B.; Hao, C.; Liu, S.; Qiu, J.; Lou, X. W. Enhancing lithium-sulphur battery performance by strongly binding the discharge products on amino-functionalized reduced graphene oxide. Nat. Commun. 2014, 5, 5002. (75) Li, G.; Sun, J.; Hou, W.; Jiang, S.; Huang, Y.; Geng, J. Threedimensional porous carbon composites containing high sulfur nanoparticle content for high-performance lithium−sulfur batteries. Nat. Commun. 2016, 7, 10601. (76) Qiu, Y.; Li, W.; Zhao, W.; Li, G.; Hou, Y.; Liu, M.; Zhou, L.; Ye, F.; Li, H.; Wei, Z.; Yang, S.; Duan, W.; Ye, Y.; Guo, J.; Zhang, Y. High-Rate, Ultralong Cycle-Life Lithium/Sulfur Batteries Enabled by Nitrogen-Doped Graphene. Nano Lett. 2014, 14, 4821−4827. (77) Wei Seh, Z.; Li, W.; Cha, J. J.; Zheng, G.; Yang, Y.; McDowell, M. T.; Hsu, P. C.; Cui, Y. Sulphur-TiO2 yolk-shell nanoarchitecture with internal void space for long-cycle lithium-sulphur batteries. Nat. Commun. 2013, 4, 1331. (78) Rehman, S.; Guo, S.; Hou, Y. Rational Design of Si/SiO2 @ Hierarchical Porous Carbon Spheres as Efficient Polysulfide Reservoirs for High-Performance Li-S Battery. Adv. Mater. 2016, 28, 3167−3172.

7084

DOI: 10.1021/acsaem.8b01498 ACS Appl. Energy Mater. 2018, 1, 7076−7084