Thiuram Vulcanization Accelerators as Polysulfide ... - ACS Publications

Jul 31, 2019 - chemical reaction between them, and the latter is quickly converted to Li2S2, along with the generation of elemental S, which will be r...
0 downloads 0 Views 2MB Size
Subscriber access provided by GUILFORD COLLEGE

Energy, Environmental, and Catalysis Applications

Thiuram vulcanization accelerators as polysulfide scavengers to suppress shuttle effects for high-performance lithium-sulfur batteries Qian Xiang, Chenyang Shi, Xueya Zhang, Lin Zhang, Liang He, Bo Hong, and Yanqing Lai ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.9b09546 • Publication Date (Web): 31 Jul 2019 Downloaded from pubs.acs.org on July 31, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Thiuram Vulcanization Accelerators as Polysulfide Scavengers to Suppress Shuttle Effects for High-Performance LithiumSulfur Batteries Qian Xiang, Chenyang Shi, Xueya Zhang, Lin Zhang, Liang He, Bo Hong* Yanqing Lai* School of Metallurgy and Environment, Central South University, Changsha Hunan 410083, China

Corresponding Authors *E-mail address: [email protected] (Yanqing Lai) [email protected] (Bo Hong)

1 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract: Lithium-sulfur (Li-S) batteries are considered one of the most promising alternatives for next-generation high-energy-density storage systems. Nevertheless, the notorious “shuttle effect” and sluggish kinetic conversion in actual operation seriously hamper its practical application. Herein, inspired by the action mechanism of vulcanization accelerators, dipentamethylenethiuram tetrasulfide (DPTT) is employed as a novel electrolyte additive. Just like a scavenger, DPTT sweeps lithium polysulfide by a spontaneous instant chemical reaction between them, and the latter is quickly converted to Li2S2, along with the generation of elemental S, which will be reduced to polysulfide again. This is beneficial for relieving the accumulation and shuttling of polysulfide in electrolyte. Therefore, Li-S batteries with DPTT-containing electrolyte exhibit enhanced capacity retention and improved rate performance. With 4 wt.% DPTT additive and 3.03-mg cm−2 S loading, the cell delivers a high initial capacity of 1227.6 mAh g−1 and excellent capacity retention of 914.7 mAh g−1 after 250 cycles at 0.5 C. This study provides fresh insight into suppressing the shuttle effect and realizing high-performance Li-S batteries.

Key words: shuttle effect, electrolyte additives, thiuram vulcanization accelerators, dipentamethylenethiuram tetrasulfide, polysulfides conversion, lithium sulfur battery.

2 ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Introduction With the rapid development of electric vehicles (EVs) and advanced portable devices, traditional lithium-ion batteries (LIBs) with limited charge-storage capacities cannot meet the booming market demand for energy-storage batteries with higher energy density.1–3 Lithium-sulfur (Li-S) batteries, with their ultrahigh theoretical energy density (2600 W h kg−1), which is far superior to state-of-the-art LIBs, have drawn great attention from researchers. Additional advantages, including low toxicity, low cost, and large abundance, make Li-S batteries promising for next-generation highenergy-density storage systems.4–7 Unfortunately, there are still several challenges that must be met to realize the practical application of Li-S batteries. The notorious “shuttle effect” triggered by soluble lithium polysulfides (LiPSs), the intrinsic low conductivity of S and its discharge product (Li2S/Li2S2), and the serious volume change of elemental S during lithiation cause rapid decay of capacity, low Coulombic efficiency, and high self-discharge rates.8–10 Recently, numerous efforts from different aspects have focused on suppressing the shuttle effect to help Li-S batteries achieve stable long cycling. Multifarious carbon materials have been designed and constructed to physically confine the dissolution and diffusion of LiPSs in bulk electrolyte.11–14 Surface modifications of the S host materials, such as doping heteroatoms and employing polar compounds, are confirmed effective approaches for trapping LiPSs and restricting shuttling.15–19 Multifunctional interlayers, binders, and coatings on the separator or cathode are introduced to block LiPS shuttling.20–23 In addition, employing functional electrolyte additives is a convenient 3 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

and efficient way to restrain the shuttling, and the widely used LiNO3 is a typical case.24–28 Nevertheless, the above-mentioned strategies all have limited effects, and the adverse impacts of LiPS shuttling cannot be completely eliminated, because even if the LiPSs are insulated in the cathode side, slow conversion reaction kinetics still result in accumulation of LiPSs during long cycles, causing increased electrolyte viscosity, reduced utilization of active S species, and even poorer contact with the conductive matrix.29,30 Consequently, batteries inevitably encounter an increase in internal resistance and fast capacity decay. Therefore, to eliminate the “shuttle effect,” it is not enough to only confine, trap, or block LiPSs in the cathode side, but introducing functional molecules to convert LiPSs efficiently is equally important. Along this line, we are devoted to identifying a functional molecule as an electrolyte addition that promotes rapid conversion of LiPSs and consequently remits its accumulation and shuttling. A vulcanization accelerator is a vital component for vulcanized rubber production. During the vulcanization process, the accelerator can easily react with S ions or radicals and speed up the cross-linking reaction with the rubber chain, which results in shortened vulcanization time and lower operation temperature. Considering the similarity between LiPSs in Li-S batteries and S ions or radicals in vulcanization, we wondered if the vulcanization accelerator can also react with LiPSs and convert them efficiently. Therefore, we researched nine different vulcanization accelerators, and the results of visualization experiments verify our idea tentatively, because six out of nine accelerators achieve varying degrees of color fading of LiPSs. By adding them to the electrolyte, we discovered that Li-S cells with thiuram 4 ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

vulcanization accelerators, especially with dipentamethylenethiuram tetrasulfide (DPTT) added, give the best performance. Based on results from ultraviolet–visiblelight (UV-Vis) absorption and other analytical methods, we propose that the intermediate products, LiPSs, react with DPTT and convert quickly to S8 and Li2S2, which significantly accelerates the conversion reaction of LiPS and restrains its diffusion and even its shuttling. Because of this, the loss of active S decreases, and the corrosion of the Li anode is relieved, and cells consequently exhibit enhanced electrochemical performance.

Experimental section Preparations of sulfur cathode. First, 10 g phenol-formaldehyde resin was added to 50 mL of ethanol to form solution A, while 2 g of cobalt nitrate hexahydrate [Co(NO3)2·6H2O], 16 g of pluronic P-123 (average Mn ~5800), and 10 g of urea (CH4N2O) were dissolved in a mixture of 10 mL of hydrochloric acid (HCl, 35–38 wt.%) and 100 mL of ethanol to form solution B. After stirring for 1 h, solutions A and B were mixed together with 25 mL of tetraethyl orthosilicate (C8H20O4Si) and stirred for another 2 h. The resulting mixture was then placed in an oven and dried at 60°C and again at 100°C for 12 h. Subsequently, the resulting rubbery product was placed in an oven for carbonization at 1100°C. Finally, the host materials were obtained by washing the carbonized product with hydrofluoric acid (AR, ≥40%) and deionized water. The S cathode was prepared by first mixing S/Co-N-C composite (70 wt.% S; Fig. S1), 10 wt.% polyvinylidene fluoride (PVDF), and 20 wt.% acetylene black in N-methyl pyrrolidone (NMP) to form a well-dispersed slurry, which was then uniformly coated on aluminum 5 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

foil and placed in a vacuum oven until the NMP volatilized completely. Unless otherwise stated, the obtained cathode has a typical mass loading of 1–1.5 mg cm -2. Preparation of electrolytes. The basic electrolyte was 1 M LiTFSI and 2 wt.% LiNO3 in DME/DOL (v:v=1:1). Nine vulcanization accelerators—tetramethylthiuram monosulfide (TMTM, 98%), tetramethylthiuram disulfide (TMTD, 98%), zinc dimethyldithiocarbamate (ZDMC, 99%), dipentamethylenethiuram disulfide (DPTD, 98.3%), dipentamethylenethiuram tetrasulfide (DPTT, 95+%), potassium ethylxanthate (PE, 98%), 2-mercaptobenzothiazole (MBT, 99%), 2,2′-dibenzothiazolyl disulfide (MBTS, 98%), and N-tert-butyl-2-benzothiazolesulfenamide (TBBS, 97%)—were purchased from Bailingwei Science and Technology Co., Ltd. (Beijing, China). The entire electrolyte configuration process was carried out in an Ar atmosphere. Synthesis of Polysulfides. Lithium sulfide(99.9%; Sigma-Aldrich) and sublimed sulfur (99.5%, Sigma-Aldrich) were added in DME/DOL(v:v=1:1) with a molar ratio of 1:5, and the mixture were magnetic stirred for 24 h at 65°C to form a dark brown solution. The overall process was operated in above-mentioned glove box. Electrochemical measurements. Electrochemical measurements were made by assembling CR2025 coin cells with Li foil as the anode. We fixed the electrolyte amount at 20 μL mg−1 S for the low S loading case (1–1.5 mg cm−2). When it was increased to 3.03 mg cm−2, 60 μL of electrolyte was added. The cells were cycled in the range of potential 1.7–2.8 V versus Li/Li+ on a LANDAN test system (Wuhan, China) after 12 h still standing. Cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) measurements were conducted on an electrochemical workstation 6 ACS Paragon Plus Environment

Page 6 of 29

Page 7 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

(Solartron 1470E). The former operated under 0.1 mV s−1 sweep rate, and the latter operated in the frequency range 0.01 Hz to 1 MHz and a voltage amplitude of 5 mV. All tests were accomplished at 25°C. Characterization. UV-Vis absorption spectra were collected with a U-4100 Spectrophotometer from 250 to 600 nm to identify changes of the polysulfide species after reacting with DPTT in electrolyte. X-ray diffraction (XRD) patterns of insoluble product after reaction were obtained with a Rigaku MiniFlex 600 using Cu Kα radiation (λ=0.154056 nm) with a scan rate of 5° min−1 from 10° to 80°. A laser microscopic confocal Raman spectrometer (Invain, Renishaw, UK) was used in the Raman testing. The surface and section morphology of the Li anode after cycling was recorded using scanning electron microscopy (SEM, Philips, FEI Quanta 200FEG), and sulfur distribution mapping images and cross-sectional line scans were collected by energydispersive spectroscopy (EDS). X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific, Inc., USA) was used to analyze chemical composition differences.

Results and discussion The constitutional formula of the nine vulcanization accelerators—tetramethylthiuram monosulfide (TMTM), tetramethylthiuram disulfide (TMTD), zinc dimethyldithiocarbamate (ZDMC), dipentamethylenethiuram disulfide (DPTD), dipentamethylenethiuram tetrasulfide (DPTT), potassium ethylxanthate (PE), 2-mercaptobenzothiazole (MBT),

2,2′-dibenzothiazolyl

disulfide

(MBTS),

and

N-tert-butyl-2-

benzothiazolesulfenamide (TBBS)—are shown in Fig. S2. To directly verify the possibility of reaction between LiPSs and accelerators, visualization experiments were 7 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

carried out. Above the row of vials depicted in Fig. 1a are a prepared 5.3 mM Li2S6 solution and the nine vulcanization accelerators with 4 wt.% addition of DME/DOL (vol:vol=1:1). Below the row of vials are solutions after reaction between Li2S6 and a corresponding vulcanization accelerator. Obviously, six of nine vials are faded, and the others exhibit a bit darker color compared with the initial Li2S6 solution. This tentatively verifies our assumption that part of the accelerators can react with LiPSs. To further investigate if accelerators really have positive impacts on Li-S batteries, we picked out six accelerators that can fade LiPSs as electrolyte additives to assemble batteries. Figure S3 shows the cycle performance of button cells with different electrolyte additives at a rate of 0.5 C (1 C=1675 mAh g−1). Except for ZDMC, the others all exhibit varying degrees of improvement in discharge capacity. Comparing the charge-discharge curves with the cells in conventional electrolyte (Fig. S4), we discovered that, different from conventional Li-S discharge curves with two apparent voltage plateaus at approximately 2.0 and 2.3 V, there is an additional slope or plateau in the high potential range, corresponding to the redox reaction of the additives.31 Overall, thiuram vulcanization accelerators, especially DPTT, show the best performance on capacity retention and Coulombic efficiency. The cycle performance of cells with conventional and DPTT-containing electrolytes is shown in Fig. 1b. The cell with 4 wt.% DPTT added delivers an initial capacity of 1266.1 mAh g−1, which is higher than 1134.2 mAh g−1 for a basic electrolyte (STD). The former retains a discharge capacity of 992.1 mAh g−1 at 0.5 C after 100 cycles, while the latter only maintains 761.3 mAh g−1. 8 ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Fig. 1 (a) Digital images of Li2S6 (5.3 mM) and nine vulcanization accelerators (4 wt.%, (1) TMTM, (2) TMTD, (3) ZDMC, (4) DPTD, (5) DPTT, (6) PE, (7) MBT, (8) MBTS, and (9) TBBS) dissolved in DOL/DME (upper row), and solutions after reaction between Li2S6 with corresponding vulcanization accelerator (lower row). (b) Cycle performance of cells with DPTT-containing and basic electrolyte. (c) Digital image showing the fast reaction between DPTT and Li2S6 within 60 s. (d) UV-Vis absorption spectra of polysulfides before and after reaction with DPTT. (e) XRD spectra of precipitate and S8 (PDF #85-0799). (f) Raman spectra comparison between precipitate and commercial S8. (g) Initial discharge curves with Li2S6 solution and supernatant after reaction with DPTT as active materials.

To further research the reason for the enhanced performance of Li-S cells with DPTT additive, we repeated the visualization experiment by adding DPTT to Li2S6 solution at a higher concentration. The color of the solution fades over time, as shown 9 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

in Fig. 1c. The original dark brown solution quickly turns pale yellow within 60 s and is accompanied by precipitation at the bottom (Fig. S5), which indicates a fast reaction between DPTT and LiPSs. To identify changes of S species after reaction, Li2S6 solutions before/after reaction are detected by a UV spectrophotometer, and UV-Vis absorption spectra are shown in Fig. 1d. The Li2S6 solution gives a discernible peak at approximately 310 nm, which represents the signal for S62− or S42−, and the shoulder peak at 420 nm is thought to be caused by S42−.32–35 The appearance of S42− is reasonable because of the inevitable disproportionation reaction.29,36,37 After reaction with DPTT, the peak at 420 nm disappears, and the peak at approximately 310 nm is significantly weakened. Note that the LiPSs are reacted, which is consistent with color fading. Instead, an emerging peak at 265 nm attributed to S22− and an unknown shoulder peak at approximately 285 nm appears.37,38 The precipitate after the reaction is collected for further detection. X-ray diffraction (Fig. 1e) and Raman analysis (Fig. 1f) both confirm the precipitate is elemental S.39,40 The Li2S6 solution and supernatant after reaction are, respectively, used as active material to assemble cells with a Li foil and mesoporous carbon wafer as a sulfur-free cathode. The initial discharge curves are shown in Fig. 1g. It can be seen that Li2S6 shows an obvious voltage plateau at ~2.1 V, which represents the reduction of low-order polysulfide (Sn2−, 2 ≤ n ≤ 4), and an inconspicuous plateau at approximately 2.3 V corresponds to the reduction of high-order polysulfide (Sn2−, 4 ≤ n ≤ 8).41 After reaction with DPTT, voltage plateaus at 2.3 and 2.1 V almost disappear and are replaced by an oblique line. Note that LiPSs are consumed after reaction with DPTT, which is consistent with the result of UV analysis, and the precipitation of active 10 ACS Paragon Plus Environment

Page 10 of 29

Page 11 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

S is responsible for the decrease of discharge capacity after reaction with DPTT. The slope under 1.9 V is related to the intrinsic property of mesoporous carbon (Fig. S6). Based on the above results and analyses, we propose that the DPTT additive could react with LiPSs via chemical reaction equations (1) and (2) below, and the product C6H10NS2Li possibly contributes to the aforementioned shoulder peak at approximately 285 nm:42,43

With the chemical conversion reactions shown above, the amount of LiPSs diffusing into the bulk electrolyte is greatly reduced during the first discharge. However, if the chemical reaction product C6H10NS2Li cannot reversibly go back to DPTT after that, the aforementioned function of DPTT could be negligible because a one-off addition of DPTT definitely cannot stop LiPS accumulation in the following cycles. The CV curves of cells with conventional and DPTT-containing electrolyte are exhibited in Fig. 2a. As shown, apart from two reduction and two oxidation peaks representing typical electrochemical reactions of S in either type of electrolyte, an additional anodic peak at ~2.6 V and a relatively cathodic peak during the anodic scan appear repetitively when DPTT is added. The additional peaks are ascribed to DPTT, which contributes a fraction of capacity, and the CV curves of DPTT-containing electrolyte with mesoporous carbon as a cathode (Fig. 2b) verify this fact.44 The chargedischarge curves (Fig. S7) and cycle performance (Fig. S8) with DPTT-containing 11 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

electrolyte manifest that DPTT indeed undergoes reversible electrochemical processes. Figure 2c and d show the charge-discharge curve at the 50th cycle with a sulfur and sulfur-free cathode in DPTT-containing electrolyte, respectively. DPTT exhibits three distinct voltage plateaus, which is consistent with CV results. Except for

Fig. 2 Cyclic voltammograms of cells (a) with conventional electrolyte and 4 wt.% DPTT additives and (b) with mesoporous carbon cathode (sulfur-free cathode) and containing-DPTT electrolyte. (c, d) the 50th charge-discharge voltage profile of cells with sulfur cathode and sulfur-free cathode in DPTTcontaining electrolyte.

the aforementioned plateau at ~2.6 V, the remaining two show the very same plateau voltages (~2.3 and ~2.05 V) as the conventional S cathode. Based on the above results and references, we cursorily speculate that DPTT experiences three primary steps 12 ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

during the discharge process shown in Eqs. (3)–(5). First, the plateau at ~2.6 V signifies the process of the DPTT being reduced to C6H10NS2Li and a double-ended radical ·S2·; the latter, which is unstable, tends to polymerize and form ·Sn· (4 ≤ n ≤ 8). The next two steps are similar to the reduction of S8, corresponding to the reduction of ·Sn· (4 ≤ n ≤ 8) and short-chain LiPSs (Li2Sn, 2 ≤ n ≤ 4).

Fig. 3 Schematic detailing the proposed mechanism of DPTT additive in Li-S batteries.

Based on the previous analysis, we confirmed that the DPTT additive is functional all through the cycling, relying on the fastly chemical conversion and reversible electrochemistry. The proposed mechanism of DPTT additive in Li-S cells is 13 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

demonstrated in Fig. 3. On the one hand, the reduction of S8 forms intermediate products, namely long-chain LiPSs (Li2Sn, 4 ≤ n ≤ 8), during the discharge process. When diffusing into electrolyte, the LiPSs react with part of DPTT Eqs. (1) and (2) immediately and are converted to Li2S2, along with the generation of S8, which will be reduced to LiPSs again. On the other hand, when the charging process is completed, the C6H10NS2Li is oxidized to DPTT again, relying on reversible electrochemical reactions Eqs. (3)–(5), and the latter will sweep part of the residual LiPSs in the electrolyte via the reactions Eqs. (1) and (2) mentioned above.42,44 Therefore, the DPTT additive helps to relieve the accumulation of LiPSs in electrolyte and effectively suppresses its shuttling.

14 ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Fig. 4 Electrochemical performance of Li-S cells with DPTT-containing electrolyte and basic electrolyte. Cycle performances (a) at 0.5 C rate, (b) at 0.5 C rate (the contributed capacity of DPTT is deducted), and (c) at 1 C rate. (d) The 1st−100th charge-discharge curves. (e) Impedance spectra. (f) Rate capability of cells with or without DPTT additive.

To further investigate and evaluate the positive effects of DPTT, a series of electrochemical tests were performed. Figure 4a displays the cyclic performance of cells with various contents of DPTT additive. The cell with basic electrolyte experiences a fast capacity fade from 1117.7 to only 618.7 mAh g−1 only after 250 15 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

cycles. When 2 wt.% DPTT was added, the cell delivered a higher initial capacity and improved capacity retention. With 4 wt.% DPTT addition, the cell showed a superior initial capacity of 1227.6 mAh g−1 and still maintains 914.7 mAh g−1 after 250 cycles. Redundant DPTT (5 wt.%) additive gave an outstanding initial capacity of 1401.6 mAh g−1 but gave an unstable cycle performance possibly because its oversaturation of electrolyte causes uncertain effects. Considering that DPTT itself has a certain capacity (Figs. S7 and S8), the effect of DPTT on battery cycle performance is unclear from Fig. 4a. Therefore, the cyclic performance comparison of cells without or with DPTT after deducting its self-contained capacity is shown in Fig. 4b. Obviously, the cell with DPTT addition experiences a much slower capacity fade. After deducting additional contribution capacity, the cell with 4 wt.% DPTT-containing electrolyte delivers 798.3 mAh g−1 after 250 cycles, which is still much higher than 618.7 mAh g−1. Better cycling performance is achieved by the cell with 4 wt.% DPTT-containing electrolyte compared with basic electrolyte when current density increased, as shown in Fig. 4c. After 150 cycles at 1 C rate, the former remains at 863.7 mAh g−1, while the latter maintains 725.5 mAh g−1. Figure 4d shows the 1st and 100th charge-discharge curves. Considering that practical Li-S batteries require higher S loading and less electrolyte, we assembled cells with an increased 4.5-mg cm−2 S loading and decreased E/S ratio (10 µL mg−1). Experimental results (Fig. S9) show that the discharge capacity of the cell with basic electrolyte shows rapid decay, especially after 20 cycles. The discharge capacity decreases to 722 mAh g−1 after 40 cycles at 0.2 C rate, while the discharge capacity with 4 wt.% DPTT-containing electrolyte maintains 944 mAh g−1. Comparison of the 16 ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

2nd and 40th charge-discharge curves with or without DPTT addition indicates that the voltage polarization shows a significant increase in basic electrolyte. Owing to the chemical reactions shown in Eqs. (1) and (2), which help sweep LiPSs in the electrolyte, the cell with DPTT addition enabled low polarization voltage and high capacity retention. When DPTT was added, an additional plateau at high potential (around 2.6 V) appeared on the discharge curves, which coincides with the results of CV curves. An obvious slope appearing on the first discharge curve after DPTT addition denotes the solid phase transition from Li2S2 to Li2S.45,46 We speculate that the reaction between LiPSs and DPTT contributes to the effective Li2S2/Li2S conversion. The electrochemical impedance spectra of cells with or without DPTT additive are displayed in Fig. 4e. The cell containing 4 wt.% DPTT-containing electrolytes before cycling possesses higher resistance probably because of increased viscosity with DPTT, but it decreases dramatically once the cycle starts and tends to stabilize after 50 cycles (Fig. S10). As a result, the resistance of cells with or without DPTT additive is almost the same after 250 cycles. Figure 4f shows the rate capability of cells without or with DPTT. For the cell without additive, the reversible capacities of 916 (third discharge), 732, 633, and 550 mAh g−1 are observed at 0.2, 0.5, 1, and 2 C, respectively. By comparison, the DPTT cell delivers higher capacities, namely 1029, 896, 809, and 709 mAh g−1 at 0.2, 0.5, 1, and 2 C, respectively. Moreover, while the rate recovers to 0.2 C, the capacity recovers to near the initial value. The improved rate capability should be ascribed to the accelerated conversion of polysulfides and additional capacity provided by DPTT.47,48 17 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5 presents the surface and cross-section morphology of Li foil after cycling. Without DPTT additive, massive unconsolidated, mossy Li deposits are found on the surface (Fig. 5a–c). In contrast, a relatively compact and uniform Li layer can be observed with DPTT-containing electrolyte (Fig. 5d–f), which means that DPTT effectively alleviates the migration of polysulfide and reduces the corrosion reaction of Li foil.49 Figure 6a–d comparatively show the morphology and EDS mapping of crosssections without or with DPTT additive after cycling. All samples are cleaned beforehand with DME solvent to remove any residue. The cycled Li foil without additive presents a rough surface morphology, riddled with cracks (Fig. 6a). Nevertheless, the latter shows a smoother and more intact surface (Fig. 6b)

Fig. 5 Surface and section morphology of Li foil after 250 cycles at 0.5 C with (a–c) basic electrolyte and (d–f) 4 wt.% DPTT-containing electrolyte.

18 ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Fig. 6 SEM images and EDS mapping of cross-section of Li foil after 100 cycles with (a,c) basic electrolyte and (b,d) 4 wt.% DPTT-containing electrolyte. (e) S 2p XPS spectra of Li foil after 100 cycles with conventional electrolyte (upper) and DPTT-containing electrolyte (lower).

and much weaker signals of S in the corresponding EDS mapping (Fig. 6d) than the former (Fig. 6c). Note that the solid electrolyte layer (SEI) formed on the cycled Li foil with DPTT additive contains less sulfurated species because of reduced corrosion reaction between Li and polysulfide.46 The EDS line scans on the cross-sections also indicate that the cycled Li foil with DPTT additive possesses lower S contents and thinner corrosion thickness (Fig. S11), which proves the remission of the shuttle effect.50,51 The XPS S 2p spectra of Li anodes without or with DPTT after 100 cycles at 0.5 C are detected to further confirm the hypothetical mechanism including accelerated conversion of polysulfides and the suppressive shuttle effect. Before the test, the surfaces of disassembled Li anodes were thoroughly cleaned with DME, and the results are presented in Fig. 6e. From the spectra of cycled Li foil without additive, one of the main chemical components of SEI is thiosulfate (168.4 and 167.2 eV), which 19 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

is considered to be the oxidation products of LiPSs by lithium nitrate (LiNO3) additive.52,53 In addition, peak binding energies of 170.2 and 169.0 eV are assigned to the S-O bond of −SO2CF3 species, which usually stems from the decomposition of LiTFSI salt.11,34,54 It is noteworthy that Li2S (161.5 and 160.3 eV) and Li2S2 (163.2 and 162.1 eV) suggest strong signal intensity, which is mainly ascribed to the reaction between Li metal and LiPSs.34,55,56 In contrast, the S 2p spectra of cycled Li foil with DPTT display much lower peaks, representing Li2S and Li2S2 and suggesting a suppressive shuttle effect.46,57 Combining the results from XPS and SEM/EDS mapping, we confirmed that the existence of DPTT accelerated the conversion of LiPSs with the reaction Eqs. (1) and (2), resulting in less LiPS diffusion to the anode side and consequent corrosion reaction. As a result, the cycled Li foil exhibits a relatively compact and uniform surface and a thinner corrosion thickness with lower S contents.

Conclusions This study found that many types of vulcanization accelerators can react with polysulfides in Li-S batteries, creating a series of interesting phenomena. A systematic study based on the best-performing dipentamethylene-thiuram tetrasulfide (DPTT) was carried out, and DPTT was found to react with soluble polysulfides instantly, promoting the formation of Li2S2 and elemental S, thus reducing the accumulation of polysulfide. Through this novel mechanism, the DPTT-containing electrolyte not only contributes extra capacity but also greatly relieves the shuttling of polysulfide in electrolyte, 20 ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

reducing the loss of active sulfur and the corrosion of Li foil. Hence, the cell with 4 wt.% DPTT additive demonstrates outstanding electrochemical performance, delivering an initial capacity of 1227.6 mAh g−1 and maintaining 914.7 mAh g−1 after 250 cycles at 0.5 C. Hundreds of vulcanization accelerators have been developed in the rubber industry, and we considered that there would be substances among them that can greatly enhance the performance of Li-S batteries. This study therefore provides a fundamental avenue for suppressing the shuttle effect and achieving high-performance Li-S batteries.

Acknowledgements This work is supported by The National Key Research and Development Program of China (2018YFB0104200).

Supporting Information The Supporting Information files are available free of charge on the ACS Publications website. TGA curve of the S/Co-N-C composite, constitutional formulas of nine vulcanization accelerators, cycle performance and charge-discharge curves of Li−S cells at 0.5 C with other vulcanization accelerators addition, cycle performance and charge-discharge curves of sulfur-free carbon cathode with 4wt% DPTT electrolyte, cycle performance of cells with/without DPTT addition at 0.2C at high S-loading and low E/S ratio operation, and cross-sectional EDS line scans (for sulfur) of the Li anode after 100 cycles in electrolyte with/without DPTT addition. 21 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References (1) Yin, Y. X.; Xin, S.; Guo, Y. G.; Wan, L. J. Lithium-sulfur batteries: electrochemistry, materials, and prospects. Angew Chem Int Ed Engl 2013, 52 (50), 13186-13200, DOI: 10.1002/anie.201304762. (2) Kang, W.; Deng, N.; Ju, J.; Li, Q.; Wu, D.; Ma, X.; Li, L.; Naebe, M.; Cheng, B. A review of recent developments in rechargeable lithium-sulfur batteries. Nanoscale 2016, 8 (37), 16541-16588, DOI: 10.1039/c6nr04923k. (3) Xiong, S.; Regula, M.; Wang, D.; Song, J. Toward Better Lithium–Sulfur Batteries: Functional Non-aqueous Liquid Electrolytes. Electrochemical Energy Reviews 2018, DOI: 10.1007/s41918-018-0015-y. (4) Lochala, J.; Liu, D.; Wu, B.; Robinson, C.; Xiao, J. Research Progress toward the Practical Applications of Lithium–Sulfur Batteries. ACS Applied Materials & Interfaces 2017, 9 (29), 24407-24421, DOI: 10.1021/acsami.7b06208. (5) 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 (48), 1606823, DOI: 10.1002/adma.201606823. (6) Ould Ely, T.; Kamzabek, D.; Chakraborty, D.; Doherty, M. F. Lithium–Sulfur Batteries: State of the Art and Future Directions. ACS Applied Energy Materials 2018, 1 (5), 1783-1814, DOI: 10.1021/acsaem.7b00153. (7) Ye, Y.; Wu, F.; Xu, S.; Qu, W.; Li, L.; Chen, R. Designing Realizable and Scalable Techniques for Practical Lithium Sulfur Batteries: A Perspective. J Phys Chem Lett 2018, 9 (6), 1398-1414, DOI: 10.1021/acs.jpclett.7b03165. (8) Manthiram, A.; Fu, Y.; Chung, S. H.; Zu, C.; Su, Y. S. Rechargeable lithium-sulfur batteries. Chem Rev 2014, 114 (23), 11751-117887, DOI: 10.1021/cr500062v. (9) Zhang, X.; Xie, H.; Kim, C. S.; Zaghib, K.; Mauger, A.; Julien, C. M. Advances in lithium—sulfur batteries. Materials Science and Engineering: R: Reports 2017, 121, 129, DOI: 10.1016/j.mser.2017.09.001. (10) He, Y.; Chang, Z.; Wu, S.; Zhou, H. Effective strategies for long-cycle life 22 ACS Paragon Plus Environment

Page 22 of 29

Page 23 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

lithium–sulfur batteries. Journal of Materials Chemistry A 2018, 6 (15), 6155-6182, DOI: 10.1039/c8ta01115j. (11) Ji, X.; Lee, K. T.; Nazar, L. F. A highly ordered nanostructured carbon-sulphur cathode for lithium-sulphur batteries. Nat Mater 2009, 8 (6), 500-506, DOI: 10.1038/nmat2460. (12) Qin, F.; Wang, X.; Zhang, K.; Fang, J.; Li, J.; Lai, Y. High areal capacity cathode and electrolyte reservoir render practical Li-S batteries. Nano Energy 2017, 38, 137146, DOI: 10.1016/j.nanoen.2017.05.037. (13) Wang, T.; Yang, Y.; Fan, L.; Wang, L.; Ma, R.; Zhang, Q.; Zhao, J.; Ge, J.; Lu, X.; Yu, X.; Yang, H.; Lu, B. Ultrathin Honeycomb-like Carbon as Sulfur Host Cathode for High Performance Lithium–Sulfur Batteries. ACS Applied Energy Materials 2018, 1 (12), 7076-7084, DOI: 10.1021/acsaem.8b01498. (14) Zhang, L.; Wang, Y.; Niu, Z.; Chen, J. Advanced nanostructured carbon-based materials for rechargeable lithium-sulfur batteries. Carbon 2019, 141, 400-416, DOI: 10.1016/j.carbon.2018.09.067. (15) See, K. A.; Jun, Y.-S.; Gerbec, J. A.; Sprafke, J. K.; Wudl, F.; Stucky, G. D.; Seshadri, R. Sulfur-Functionalized Mesoporous Carbons as Sulfur Hosts in Li–S Batteries: Increasing the Affinity of Polysulfide Intermediates to Enhance Performance. ACS Applied Materials & Interfaces 2014, 6 (14), 10908-10916, DOI: 10.1021/am405025n. (16) Peng, Y.; Zhang, Y.; Huang, J.; Wang, Y.; Li, H.; Hwang, B. J.; Zhao, J. Nitrogen and oxygen dual-doped hollow carbon nanospheres derived from catechol/polyamine as sulfur hosts for advanced lithium sulfur batteries. Carbon 2017, 124, 23-33, DOI: 10.1016/j.carbon.2017.08.035. (17) Liu, J.; Yuan, L.; Yuan, K.; Li, Z.; Hao, Z.; Xiang, J.; Huang, Y. SnO2 as a highefficiency polysulfide trap in lithium-sulfur batteries. Nanoscale 2016, 8 (28), 1363813645, DOI: 10.1039/c6nr02345b. (18) Zhou., G.; Tian., H.; Jin., Y.; Tao., X.; Liu., B.; Zhang., R.; Seh., Z. W.; Zhuo., D.; Liu., Y.; Sun., J.; Zhao., J.; Zu., C.; Wu., D. S.; Zhang., Q.; Cui., Y. Catalytic oxidation 23 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 29

of Li2S on the surface of metal sulfides for Li-S batteries. PNAS 2017, 114 (5), 840845. (19) Li, L.; Chen, L.; Mukherjee, S.; Gao, J.; Sun, H.; Liu, Z.; Ma, X.; Gupta, T.; Singh, C. V.; Ren, W.; Cheng, H.-M.; Koratkar, N. Phosphorene as a Polysulfide Immobilizer and Catalyst in High-Performance Lithium-Sulfur Batteries. Advanced Materials 2017, 29 (2), 1602734, DOI: 10.1002/adma.201602734. (20) Huang, J.-Q.; Zhang, Q.; Wei, F. Multi-functional separator/interlayer system for high-stable lithium-sulfur batteries: Progress and prospects. Energy Storage Materials 2015, 1, 127-145, DOI: 10.1016/j.ensm.2015.09.008. (21) Fan, W.; Zhang, L.; Liu, T. Multifunctional second barrier layers for lithium–sulfur batteries.

Materials

Chemistry

Frontiers

2018,

2

(2),

235-252,

DOI:

10.1039/c7qm00405b. (22) Liu, T.; Sun, S.; Hao, J.; Song, W.; Niu, Q.; Sun, X.; Wu, Y.; Song, D.; Wu, J. Reliable Interlayer Based on Hybrid Nanocomposites and Carbon Nanotubes for Lithium–Sulfur Batteries. ACS Applied Materials & Interfaces 2019, 11 (17), 1560715615, DOI: 10.1021/acsami.9b02136. (23) Zhou, G.; Liu, K.; Fan, Y.; Yuan, M.; Liu, B.; Liu, W.; Shi, F.; Liu, Y.; Chen, W.; Lopez, J.; Zhuo, D.; Zhao, J.; Tsao, Y.; Huang, X.; Zhang, Q.; Cui, Y. An Aqueous Inorganic Polymer Binder for High Performance Lithium–Sulfur Batteries with FlameRetardant Properties. ACS Central Science 2018, 4 (2), 260-267, DOI: 10.1021/acscentsci.7b00569. (24) Aurbach, D.; Pollak, E.; Elazari, R.; Salitra, G.; Kelley, C. S.; Affinito, J. On the Surface Chemical Aspects of Very High Energy Density, Rechargeable Li–Sulfur Batteries. Journal of The Electrochemical Society 2009, 156 (8), A694-A702, DOI: 10.1149/1.3148721. (25) Scheers, J.; Fantini, S.; Johansson, P. A review of electrolytes for lithium–sulphur batteries.

Journal

of

Power

Sources

2014,

255,

204-218,

DOI:

10.1016/j.jpowsour.2014.01.023. (26) Gu, S.; Wen, Z.; Qian, R.; Jin, J.; Wang, Q.; Wu, M.; Zhuo, S. Carbon Disulfide 24 ACS Paragon Plus Environment

Page 25 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Cosolvent Electrolytes for High-Performance Lithium Sulfur Batteries. ACS Appl Mater Interfaces 2016, 8 (50), 34379-34386, DOI: 10.1021/acsami.6b11619. (27) Yang, W.; Yang, W.; Song, A.; Gao, L.; Sun, G.; Shao, G. Pyrrole as a promising electrolyte additive to trap polysulfides for lithium-sulfur batteries. Journal of Power Sources 2017, 348, 175-182, DOI: 10.1016/j.jpowsour.2017.03.008. (28) Zhang, L.; Ling, M.; Feng, J.; Mai, L.; Liu, G.; Guo, J. The synergetic interaction between LiNO3 and lithium polysulfides for suppressing shuttle effect of lithium-sulfur batteries. Energy Storage Materials 2018, 11, 24-29, DOI: 10.1016/j.ensm.2017.09.001. (29) Chen., Y.; Zhang., H.; Xu., W.; Yang., X.; Yu., Y.; Li., X.; Zhang., H. Polysulfide Stabilization: A Pivotal Strategy to Achieve High Energy Density Li-S Batteries with Long Cycle Life. Advanced Function Materials 2017, 28 (8), 1704987, DOI: 10.1002/adfm.201704987. (30) Shi., H.; Lv., W.; Zhang., C.; Wang., D.-W.; Ling., G.; He., Y.; Kang., F.; Yang., Q.-H. Functional Carbons Remedy the Shuttling of Polysulfides in Lithium-Sulfur Batteries: Confining, Trapping, Blocking,and Breaking up. Advanced Functional Materials 2018, 28 (38), 1800508, DOI: 10.1002/adfm.201800508. (31) Song, J.; Noh, H.; Lee, H.; Lee, J.-N.; Lee, D. J.; Lee, Y.; Kim, C. H.; Lee, Y. M.; Park, J.-K.; Kim, H.-T. Polysulfide rejection layer from alpha-lipoic acid for high performance lithium–sulfur battery. Journal of Materials Chemistry A 2015, 3 (1), 323330, DOI: 10.1039/c4ta03625e. (32) Barchasz, C.; Molton, F.; Duboc, C.; Lepretre, J. C.; Patoux, S.; Alloin, F. Lithium/sulfur cell discharge mechanism: an original approach for intermediate species identification. Anal Chem 2012, 84 (9), 3973-3980, DOI: 10.1021/ac2032244. (33) Hua, W.; Yang, Z.; Nie, H.; Li, Z.; Yang, J.; Guo, Z.; Ruan, C.; Chen, X.; Huang, S. Polysulfide-Scission Reagents for the Suppression of the Shuttle Effect in LithiumSulfur Batteries. ACS Nano 2017, 11 (2), 2209-2218, DOI: 10.1021/acsnano.6b08627. (34) Zhang, G.; Zhang, Z.-W.; Peng, H.-J.; Huang, J.-Q.; Zhang, Q. A Toolbox for Lithium-Sulfur Battery Research: Methods and Protocols. Small Methods 2017, 1 (7), 1700134, DOI: 10.1002/smtd.201700134. 25 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(35) Zou, Q.; Lu, Y. C. Solvent-Dictated Lithium Sulfur Redox Reactions: An Operando UV-vis Spectroscopic Study. J Phys Chem Lett 2016, 7 (8), 1518-1525, DOI: 10.1021/acs.jpclett.6b00228. (36) Jiang, G.; Xu, F.; Yang, S.; Wu, J.; Wei, B.; Wang, H. Mesoporous, conductive molybdenum nitride as efficient sulfur hosts for high-performance lithium-sulfur batteries. Journal of Power Sources 2018, 395, 77-84, DOI: 10.1016/j.jpowsour. 2018.05.061. (37) Klein, M. J.; Goossens, K.; ChristopherW; Bielawski; Manthiram, A. Elucidating the Electrochemical Activity of Electrolyte-Insoluble Polysulfide Species in LithiumSulfur Batteries. Journal of The Electrochemical Society 2016, 163 (9), A2109-A2116, DOI: 10.1149/2.0051610jes]. (38) Song, Y.; Zhao, W.; Zhu, X.; Zhang, L.; Li, Q.; Ding, F.; Liu, Z.; Sun, J. Vanadium Dioxide-Graphene Composite with Ultrafast Anchoring Behavior of Polysulfides for Lithium-Sulfur Batteries. ACS Appl Mater Interfaces 2018, 10 (18), 15733-15741, DOI: 10.1021/acsami.8b02920. (39) Wu, R.; Chen, S.; Deng, J.; Huang, X.; Song, Y.; Gan, R.; Wan, X.; Wei, Z. Hierarchically porous nitrogen-doped carbon as cathode for lithium-sulfur batteries. Journal of Energy Chemistry 2018, 27 (6), 1661-1667, DOI: 10.1016/j.jechem. 2018.02.010. (40) Hagen, M.; Schiffels, P.; Hammer, M. In-Situ Raman Investigation of Polysulfide Formation in Li-S Cells. Journal of The Electrochemical Society 2013, 160 (8), A1205A1214, DOI: 10.1149/2.045308jes. (41) Li, L.; Wu, Z. P.; Sun, H.; Chen, D.; Gao, J.; Suresh, S.; Chow, P.; Singh, C. V.; Koratkar, N. A Foldable Lithium-Sulfur Battery. ACS Nano 2015, 9 (11), 11342-11350, DOI: 10.1021/acsnano.5b05068. (42) Liu, M.; Visco, S. J.; C, L.; Jonghe, D. Electrochemical Properties of Organic Disulfide/Thiolate Redox Couples. Journal of The Electrochemical Society 1989, 136 (9), 2570-2575. (43) Visco, S. J.; Liu, M.; Jonghe, L. C. D. Ambient Temperature High-Rate 26 ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Lithium/Organosulfur Batteries. 1990, 137 (4), 1191-1192. (44) Bhargav, A.; Ma, Y.; Shashikala, K.; Cui, Y.; Losovyj, Y.; Fu, Y. The unique chemistry of thiuram polysulfides enables energy dense lithium batteries. Journal of Materials Chemistry A 2017, 5 (47), 25005-25013, DOI: 10.1039/c7ta07460c. (45) Lin, Z.; Liang, C. Lithium–sulfur batteries: from liquid to solid cells. Journal of Materials Chemistry A 2015, 3 (3), 936-958, DOI: 10.1039/c4ta04727c. (46) Lu, Y.; Gu, S.; Guo, J.; Rui, K.; Chen, C.; Zhang, S.; Jin, J.; Yang, J.; Wen, Z. Sulfonic Groups Originated Dual-Functional Interlayer for High Performance LithiumSulfur Battery. ACS Appl Mater Interfaces 2017, 9 (17), 14878-14888, DOI: 10.1021/acsami.7b02142. (47) Zhang, L.; Chen, X.; Wan, F.; Niu, Z.; Wang, Y.; Zhang, Q.; Chen, J. Enhanced Electrochemical Kinetics and Polysulfide Traps of Indium Nitride for Highly Stable Lithium-Sulfur

Batteries.

ACS

Nano

2018,

12

(9),

9578-9586,

DOI:

10.1021/acsnano.8b05466. (48) Chen, S.; Dai, F.; Gordin, M. L.; Yu, Z.; Gao, Y.; Song, J.; Wang, D. Functional Organosulfide Electrolyte Promotes an Alternate Reaction Pathway to Achieve High Performance in Lithium-Sulfur Batteries. Angew Chem Int Ed Engl 2016, 55 (13), 4303-4307, DOI: 10.1002/anie.201511830. (49) Ma, G.; Wen, Z.; Wu, M.; Shen, C.; Wang, Q.; Jin, J.; Wu, X. A lithium anode protection guided highly-stable lithium-sulfur battery. Chem Commun (Camb) 2014, 50 (91), 14209-14212, DOI: 10.1039/c4cc05535g. (50) Zu, C.; Manthiram, A. Stabilized Lithium-Metal Surface in a Polysulfide-Rich Environment of Lithium-Sulfur Batteries. J Phys Chem Lett 2014, 5 (15), 2522-2527, DOI: 10.1021/jz501352e. (51) Wang, P.; Zhang, Z.; Yan, X.; Xu, M.; Chen, Y.; Li, J.; Li, J.; Zhang, K.; Lai, Y. Pomegranate-like microclusters organized by ultrafine Co nanoparticles@nitrogendoped carbon subunits as sulfur hosts for long-life lithium-sulfur batteries. Journal of Materials Chemistry A 2018, 6 (29), 14178-14187, DOI: 10.1039/c8ta04214d. (52) Zhao, T.; Ye, Y.; Peng, X.; Divitini, G.; Kim, H.-K.; Lao, C.-Y.; Coxon, P. R.; Xi, 27 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 29

K.; Liu, Y.; Ducati, C.; Chen, R.; Kumar, R. V. Advanced Lithium-Sulfur Batteries Enabled by a Bio-Inspired Polysulfide Adsorptive Brush. Advanced Functional Materials 2016, 26 (46), 8418-8426, DOI: 10.1002/adfm.201604069. (53) Liang, X.; Kwok, C. Y.; Lodi-Marzano, F.; Pang, Q.; Cuisinier, M.; Huang, H.; Hart, C. J.; Houtarde, D.; Kaup, K.; Sommer, H.; Brezesinski, T.; Janek, J.; Nazar, L. F. Tuning Transition Metal Oxide-Sulfur Interactions for Long Life Lithium Sulfur Batteries: The “Goldilocks” Principle. Advanced Energy Materials 2016, 6 (6), 1501636, DOI: 10.1002/aenm.201501636. (54) Yang, Y. B.; Liu, Y. X.; Song, Z.; Zhou, Y. H.; Zhan, H. Li(+)-Permeable Film on Lithium Anode for Lithium Sulfur Battery. ACS Appl Mater Interfaces 2017, 9 (44), 38950-38958, DOI: 10.1021/acsami.7b10306. (55) Mosavati, N.; Salley, S. O.; Ng, K. Y. S. Characterization and electrochemical activities of nanostructured transition metal nitrides as cathode materials for lithium sulfur

batteries.

Journal

of

Power

Sources

2017,

340,

210-216,

DOI:

10.1016/j.jpowsour.2016.11.033. (56) Agostini, M.; Xiong, S.; Matic, A.; Hassoun, J. Polysulfide-containing Glymebased Electrolytes for Lithium Sulfur Battery. Chemistry of Materials 2015, 27 (13), 4604-4611, DOI: 10.1021/acs.chemmater.5b00896. (57) Shen, X.; Qian, T.; Chen, P.; Liu, J.; Wang, M.; Yan, C. Bioinspired Polysulfiphobic Artificial Interphase Layer on Lithium Metal Anodes for Lithium Sulfur Batteries. ACS Appl Mater Interfaces 2018, 10 (36), 30058-30064, DOI: 10.1021/acsami.8b12093.

28 ACS Paragon Plus Environment

Page 29 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

TOC graphic for manuscript

29 ACS Paragon Plus Environment