Heterogeneous Catalysis for Lithium–Sulfur ... - ACS Publications

Jan 5, 2017 - ACS Energy Lett. , 2017, 2 (2), pp 327–333 ... Nitride Hollow Spheres as an Efficient SeS 2 Host for Advanced Li-SeS 2 Batteries. Zhen...
0 downloads 0 Views 1MB Size
Subscriber access provided by MONASH UNIVERSITY

Letter

Heterogeneous Catalysis for Lithium-Sulfur Batteries: Enhanced Rate Performance by Promoting Polysulfide Fragmentations Tae-Gyung Jeong, Dong Shin Choi, Hannah Song, Jihwan Choi, ShinAe Park, Si Hyoung Oh, Heejin Kim, Yousung Jung, and Yong-Tae Kim ACS Energy Lett., Just Accepted Manuscript • DOI: 10.1021/acsenergylett.6b00603 • Publication Date (Web): 05 Jan 2017 Downloaded from http://pubs.acs.org on January 9, 2017

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 free 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 accessible to all readers and 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.

ACS Energy Letters 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 23

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 Energy Letters

Heterogeneous Catalysis for Lithium-Sulfur Batteries: Enhanced Rate Performance by Promoting Polysulfide Fragmentations Tae-Gyung Jeong,a† Dong Shin Choi,b† Hannah Song,a† Jihwan Choi,a Shin-Ae Park,a Si Hyoung Oh,d Heejin Kim,*c Yousung Jung,*b Yong-Tae Kim*a a.

Department of Energy Systems, Pusan National University, Busan 46241, Republic of Korea.

b.

Graduate School of EEWS, Korea Advanced institute of Science and Technology, Daejeon

34141, Republic of Korea. c.

Center for Electron Microscopy Research, Korea Basic Science Institute, Suncheon 57922,

Republic of Korea. d.

Centre for Energy Convergence, Korea Institute of Science and Technology, Seoul 02792,

Republic of Korea. Corresponding Author *E-mail: [email protected] (Y. K) *Email: [email protected] (Y. J) *Email: [email protected] (H. K)

ACS Paragon Plus Environment

1

ACS Energy Letters

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 2 of 23

Abstract

A spatial confiment of polysulfides using the metal compound additives having polar surfaces has been considered to be a promising approach to address the insufficient rate capability and cyclability of lithium-sulfur batteries. Herein, we report more effective approach outperforming this conventional one; a heterogeneous catalysis to promote polysulfide fragmentations. It was revealed using the combined computational and experimental approaches that an ultra-strong adsorption of elemental sulfur on TiN surfaces resulted in a spontaenous fragmentation into shorter chain of polysulfides. This heterogeneous catalysis reaction improved the sluggish kinetics of polysulfide reduction due to the chemical disproportionation at the second plateau. A markedly enhanced rate capability was finally obtained, exhibiting a discharge capacity of 700 mAh g-1 at a scan rate of 5C.

TOC GRAPHICS

ACS Paragon Plus Environment

2

Page 3 of 23

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 Energy Letters

Lithium-sulfur batteries (LSBs) have attracted a significant recent attention due to a high theoretical capacity of 1675 mAh g-1 and specific energy of 2600 Wh kg-1 compared to the conventional intercalation-based lithium-ion batteries (LIBs). Sulfur is a naturally abundant element and an inexpensive material, making it even more attractive for large scale energy storage applications.1-5 The main drawbacks in commercializing the LSBs, however, are known to be a low cycle life and poor rate capability, originating partly from the insulating nature of active materials and from the dissolution of polysulfides into the electrolytes during the cell operations.6-8 In an effort to overcome these issues, exciting progresses have been made in the cathodes side by developing new structures and composites to prevent the loss of active materials due to dissolution as well as to improve the conductivity of the cathode at the same time. To physically trap the polysulfides, carbon hosts with high porosity,9-12 carbon interlayers,13-15 hollow carbon nanostructures16-18 , graphene-based structures19-21 and polysulfide absorbent addition22 have been employed due to their high electrical conductivity. While these efforts have increased the capacities and cyclabilities of the existing sulfur cathodes to some extent, nearly 20% of the active material is still lost in the electrolytes.9 Especially at a reasonably high current density, not all the sulfur molecules participate in reduction reactions during discharge but rather remains as an elemental form, leading to a low capacity.23 Recently, chemical modifications of the electrode surfaces with oxides, sulphides, nitrides or polymers have been applied to retain the polysulfides near the cathode.24-30 Also, the use of metal/metal oxides modified carbon to adsorb and re-activate polysulfide species has exhibited excellent capacity and cycle performances.31-34 The addition of these materials provides the polar surface that can strongly interact with the polysulfides and spatially confine them near the

ACS Paragon Plus Environment

3

ACS Energy Letters

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 4 of 23

electrode surfaces, addressing the shuttle effects more aggressively. Most of previous studies have explained the effects of such additives mainly by relying on their increased binding affinities to the polysulfides although they may have more complex and richer functionalities and thus can affect the battery performance via different aspects. Herein, we show that a heterogeneous catalysis on metal compound surfaces having an ultra-strong bond formation with elemental sulfur like nitrogen-terminated TiN results in a spontaneous fragmentation of sulfur into shorter chain of polysulfides. In particular, we demonstrate that this unreported phenomenon can markedly enhance the rate capability with accelerated feeding of short polysulfides by the facilitated fragmentation of longer polysulfides particularly at about 2.0 V region corresponding to the second plateau in the charge-discharge curves, where the further reduction is restricted by the insufficient supply of polysulfides to the electrode due to the sluggish chemical disproportionation.35 We begin with the demonstration of battery performance changes with various additives in sulfur-carbon composite cathodes before discussing the heterogeneous catalysis in LSB in detail. We prepared four composite cathodes for LSB with and without Ti-based additives: 1) bare sulfur–carbon mixture (bare), 2) sulfur–carbon mixture with TiO2 additive (C/TiO2), 3) sulfur– carbon mixture with Ti4O7 additive (C/Ti4O7), and 4) sulfur–carbon mixture with TiN additive (C/TiN). All components in each cathode were homogenously mixed as evidenced by the EPMA images (Figure S1 in the ESI). Also, their morphology was nearly identical as shown in Figure S2 in the ESI, suggesting that the influences from the particle morphology between samples could be neglected. The electrochemical properties of the C/TiN cathode were compared with the other composite cathodes via the conventional galvanostatic cycling. Figure 1 shows the voltage profile of LSBs

ACS Paragon Plus Environment

4

Page 5 of 23

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 Energy Letters

Figure 1. Charge-discharge profiles at (a) 0.2C, (b) 0.5C, (c) 1C, (d) 2C, (e) 5C, and (f) discharge capacity ratio of C/TiN to bare electrodes. (g) Cyclability of the bare, C/TiN, C/Ti4O7 and C/TiO2 electrodes.

ACS Paragon Plus Environment

5

ACS Energy Letters

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 6 of 23

on different C-rates varying from 0.2 C to 5 C. In every C-rate, all the LSBs present typical voltage profiles as reported in literatures, which include a rather short plateau at 2.4 V and a long plateau at 2.0 V. Among the Ti-based additives considered here, TiN markedly improves the battery properties in all aspects, such as discharge capacity, cycle performance, rate capability and polarization. The discharge capacities of the bare, C/TiO2, C/Ti4O7, and C/TiN cathodes were 724, 954, 946, and 1069 mAh g-1 at a scan rate of 0.2 C (Figure 1a), and 231, 294, 288, and 579 mAh g-1 at a scan rate of 2 C (Figure 1d), respectively. As proposed in previous works, the increased capacity for the latter samples can be attributed to the strong interaction between the polysulfide and polar surface of the Ti-based additives.27,36 The C/TiN cathode also outperforms the other LSBs in the cycle retention as shown in Figure 1f. After 50 cycles, the C/TiN cathode still retains 70% of the initial discharge capacity, while the bare, C/TiO2 and C/Ti4O7 cathodes respectively exhibit only 55%, 28% and 47% of each initial capacity after the same cycle. Furthermore, the C/TiN cathode exhibits a superior rate capability, achieving the discharge capacity of 780 mAh g-1 at a rate of 1 C; and a capacity of 411 mAh g-1 is preserved even at a higher scan rate of 5 C. On the other hand, C/TiO2 and C/Ti4O7 cathodes lose their functionalities at fast discharge conditions. See Figure S3 in the ESI for more detailed comparison between Tibased additives. In this set of measurements, it is noteworthy that the capacity improvement by TiN additive is more significant at a faster cell operation (Figure 1g). The discharge capacity ratio of the C/TiN to bare cathode, namely the effect of TiN, increases from 1.5 at a scan rate of 1 C to 5.5 at a scan rate of 5 C. It is also remarkable that the C/TiN cathode shows smaller polarizations than the other samples at high C-rates, especially on the second plateau. These results all point to demonstrate that the TiN additive improves the cell kinetics more effectively than the other additives.

ACS Paragon Plus Environment

6

Page 7 of 23

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 Energy Letters

Figure 2. Battery performance enhancement with TiN additives. (a) Capacity enhancement in the first and second plateaus on different c-rate, compared to the bare cathode. (b) Polarization for the second plateau in bare and C/TiN cathodes, in reference to the values at 0.2 C. (c) CV curves with and without TiN. For further details, we compared the variances in discharge capacity and polarization on the first (2.4 V) and second (2.0 V) plateaus separately. As shown in Figure 2, both discharge capacity and polarization are improved more significantly on the second plateau with adding TiN, suggesting that the TiN additive contributes mainly to this 2 V region. This result is consistent with the previous studies that the later parts of the discharge processes, corresponding to the

ACS Paragon Plus Environment

7

ACS Energy Letters

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 8 of 23

second plateau, is relatively slower and is more influenced by the kinetic factors compared to the first plateau.37,38 The CV measurements of Figure 2c also clearly show that the enhancement by TiN is notable for the second plateau (2.03 V at negative sweep) compared to the first plateau (2.4 V at negative sweep). Those for the other cells (S-TiO2, S-Ti4O7) are shown in Figure S4. Several aspects have been suggested as the origin of slow reaction kinetics at the second plateau region, such as insoluble and insulating nature of the Li2S products,39,40 slow rates of disproportionation reactions,41 and the shuttling effects.42,43 Thus, considering the remarkable improvements shown for the second plateau behavior simply by adding TiN, it is conceivable that the TiN additive may improve these existing issues of slow kinetics of the sulfur battery by, for example, helping to form Li2S, facilitating the disproportionation of long polysulfides, and anchoring polysulfides to prevent the shuttle effect. To gain further insights into the role of TiN additives along these mechanisms, we investigated the interaction between calculations. As shown in Figure 3a, the binding energy of a single S8 molecule on the TiN surface is 6.60 eV, which is substantially greater than that on the graphitic (0.76 eV), TiO2 (1.78 eV), and Ti4O7 (1.28 eV) surfaces (binding geometries are presented in Figure S6 in the ESI). This binding strength of 6.60 eV is also considerably larger than the dispersion interactions appearing in other sulfide or oxide surfaces of the previous work (~0.8 eV).44 This unprecedentedly strong binding of S to TiN lies in the fact that all Ti atoms on the reconstructed TiN(001) surface are coordinatively unsaturated, leading to the complete collapsing of S8 molecules into the atomic sulfur by forming Ti–S bonds to stabilize the unfavorable coordination. The ab initio molecular dynamics simulation (See ESI) consistently predicts the passivation of TiN surface with S atoms as much as ~1 monolayer as shown in Figure 3b. Note that both C and Li atoms, which are the most abundant elements in this

ACS Paragon Plus Environment

8

Page 9 of 23

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 Energy Letters

Figure 3. (a) Binding energy of S8 molecule on different Ti-based surfaces. (b) Sulfur passivated TiN (S–TiN) surface. (c) XPS S 2p spectra of elemental sulfur and (d) S8 on TiN. (e) Formation energy of LiPS with and without S-TiN, and their difference. (f) Sulfur reduction reaction on TiN (g) Binding geometry of Li2S4 on the S–TiN surface.

ACS Paragon Plus Environment

9

ACS Energy Letters

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 10 of 23

this composite cathode besides S, are unstable on the TiN surface, allowing the Ti-S bond primarily. We confirmed the existence of surficial Ti–S bonds using the XPS measurements. As shown in Figure 3c and 3d, the binding energy of the S–2p electron shifts from 164.09 eV in the elemental S8 to 163.10 eV in the toluene-washed TiN/S composite without any trace peak of S8. This red shift of the peak can be ascribed to the increased electron density around the S atoms, as in the various metal sulfides.45 The coincident shifts of calculated charge density (0, +0.16e, and +0.66e for S8, S-TiN, and TiS2, respectively) and experimental XPS data (164.09, 163.10, and 161.9 eV for S8, S-TiN, and TiS2, respectively) suggests that the surface-bound S8 rings are completely disrupted and all S atoms make the chemical bond with Ti. We note that S prefers to form a bond with Ti than N on the TiN surface by 0.51 eV. This sulfur passivated TiN (which we denote as S–TiN) surface shows a favorable interaction with both lithium polysulfides (LiPS) and a sulfur molecule. It has been proposed that the anchoring of LiPS on electrode materials is most effective when they exhibit a moderate binding affinity to the surface, as for sulfides.44 The S–TiN surface shows a comparable (albeit slightly smaller) binding affinity to the LiPS when compared to the TiS2 surface, one of the promising anchoring materials with optimal binding strengths.44 (see Figure S7) The S–TiN surface also shows the same characteristic trend in binding affinity as the other moderate anchoring materials; that is, the binding affinity for shorter LiPS increases overall while the affinity for Li2S6 slightly decreases. The structural similarity of the LiPS conformations on the previous TiS2 and the present S–TiN surfaces is another clear indication for stable anchoring of active materials on the S–TiN (Figure S8 in the ESI) surface.

ACS Paragon Plus Environment

10

Page 11 of 23

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 Energy Letters

Figure 3e shows the formation energy of LiPS when they are in a molecular form (Efmolecule) and when they are bound on the S–TiN surface (EfS-TiN), in reference to the lithium metal and alpha sulfur phases, defined as below. Efmolecule (Li2Sn) = E(Li2Sn) – 2 E(Limetal) – n/8 E(S8)

(1)

EfS-TiN (Li2Sn) = E(Li2Sn@S-TiN) – E(S-TiN) – 2E(Limetal) – n/8 E(S8)

(2)

In the molecular form without the S–TiN surface, the long LiPS (n ≥ 4 in Li2Sn) are more stable than the short LiPS (n ≤ 2 in Li2Sn), impeding the formation of short LiPS. Recently, Wang et al. demonstrated that the clustering of LiPS stabilizes the short LiPS more drastically than the long LiPS,46 leading to the desired precipitation of Li2S2 and Li2S. In our calculations, the Li2S cluster, denoted as [Li2S]m, becomes more stable than the long chain LiPS molecules when m > 4, which can be considered as a critical size for the crystallization (details are in Figure S9 in the ESI). In the presence of the S–TiN surface, however, this stability trend of LiPS is reversed, that is, the short LiPS are rather more stable than the long LiPS even without clustering. It is due to the fact that all atoms in the short LiPS form strong chemical bonds with the surface, while the long LiPS interacts with the surface via relatively weak dispersion forces. From equation (1) and (2), the formation energy difference with and without S-TiN surface can be arranged as a follow: EfS-TiN (Li2Sn) – Efmolecule (Li2Sn) = E(Li2Sn@S-TiN) – E(S-TiN) – E(Li2Sn) where the term on the right-hand side corresponds to the binding energy of LiPS on the S-TiN surface. The difference plot, i.e., binding energy of LiPS, show a clear effect of S-TiN surface as drawn with a dotted line in Figure 3e, where the short LiPS are stabilized more than the long

ACS Paragon Plus Environment

11

ACS Energy Letters

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 12 of 23

LiPS. This peculiar stabilization of the short LiPS on the S–TiN surface promotes the later part of the discharge process that is related to the fragmentation into short LiPS. Figure 3f is the CV curves for the pure polycrystalline TiN thin film formed on bulk Ti pellet without any conductor or binder. The TiN itself does not show any electrochemical activity toward Li, consistent with the aforementioned result that the Li does not bind on the TiN surface. Contrarily, when 4mM of S8 is added to the electrolyte, the TiN cathode exhibits a small but sharp peak at 1.5–1.8 V vs. Li/Li+ in a negative (discharge) sweep. This voltage value agrees with the calculated lithiation voltage of the S–TiN surface (1.70 V vs. Li/Li+), implying that the reduction reaction can readily be proceeded on the S–TiN surface due to a high electronic conductivity of TiN (>103 S cm-1 for bulk).42 In other words, the role of TiN is not only promoting the fragmentation of LiPS, but also participating in the electrochemical reactions, in contrast to the other insulating anchoring materials. Nevertheless, the electronic conductivity of Ti4O7 (~103 S cm-1 for bulk)48,49 is also sufficient for sustaining the electron transfer, yet still showed a much smaller enhancement in LSB performance compared to the C/TiN electrode (Figure 1), suggesting that the high electronic conductivity of additives can be helpful but not the primary factor that enhances the LBS performance presently observed. The improved rate capability of TiN compared to the Ti4O7 is clearly shown in Figure S3 in the ESI. Since the second plateau is related to the short LiPS (n ≤ 4 in Li2Sn),37,38 we further investigated the binding of Li2S4 and its fragmentation on the S–TiN surface. As shown in Figure 3g, the Li2S4 adheres to the S-TiN surface via a side-on mode as in the case of other sulfides. When forming these bonds, approximately 0.65 e- (from S of Li2S4) transfer to the S–TiN substrate (by 0.31 e-) and to the Li of Li2S4 (by 0.34 e-), weakening the intramolecular bond strengths of Li2S4 when adsorbed on the S-TiN surface. The Li–S and S–S bond lengths in the S-

ACS Paragon Plus Environment

12

Page 13 of 23

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 Energy Letters

TiN-adsorbed Li2S4 are also elongated by 0.15 Å and 0.03 Å, respectively, compared to those in the unbound Li2S4. This destabilization of Li2S4 on the S-TiN surface then promotes the further fragmentation of this polysulfide to shorter LiPS with an ultra-strong adsorption which can be considered to be a kind of heterogeneous catalysis. It is obvious that this process is very helpful for the subsequent reduction reaction and/or disproportionation reactions. For example, the reduction reaction, Li2S4 + 2Li + 2e-  2Li2S2, is endothermic (nonspontaneous) by 0.41 eV without the substrate, but it becomes exothermic (spontaneous) by 2.33 eV on the S–TiN surface. Also, although the disproportionation reaction such as Li2S4  Li2S2 + 2/8S8, for example, is well known to be a source of slow kinetics for the second plateau and indeed is calculated here to be energetically uphill (endothermic) by 0.87 eV without the help of the additives, the same reaction on the S–TiN surface becomes significantly less difficult, now requiring only 0.24 eV on the S–TiN surface. On the basis of these thermodynamic driving forces, the activation barrier for fragmentations that correspond to the intramolecular bond breaking of Li2S4 are also expected to be lower on the S-TiN surface, improving the reaction kinetics, as shown experimentally here. The interesting role of TiN as a promoter of chemical disproportionation was clearly confirmed with in-situ UV-Vis spectroscopy. In order to compare the rate of chemical disproportionation on the surface of carbon and TiN, glassy carbon and TiN bulk surfaces were immersed into the electrolyte containing chemically synthesized 30 mM Li2S4. As shown in Figure S5, a strong absorption peak corresponding to the Li2S4 was initially demonstrated at around 410 nm and the peak was decreased with time by progress of chemical disproportionation. Interestingly, the rate of peak decrease is much faster for TiN than for glassy carbon, indicating that TiN can promote the chemical disproportionation.

ACS Paragon Plus Environment

13

ACS Energy Letters

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 14 of 23

Lastly, to verify compatibility of the TiN additive to other polysulfide capturing techniques, we additionally inserted acetylene black mesh (ABM) between the cathode and separator as demonstrated in the previous work.15 Figure 4 shows a comparison of LSB cathodes for the bare, C/TiN, ABM-inserted, and ABM-inserted C/TiN (C/TiN–ABM) cases. The order of capacity

Figure 4. Rate capability of bare cathode, C/TiN cathode, cathode with a carbon interlayer (ABM), and C/TiN cathode with ABM (C/TiN–ABM). *ABM from reference 15. retention with varying the scan rate is C/TiN–ABM > C/TiN > ABM > bare cathodes, where the C/TiN–ABM LSBs exhibits an improved discharge capacity by a factor of 10 compared to the bare cathode at a scan rate of 5C (700 mAh g-1 vs. 70 mAh g-1). This clearly shows that the combination of physical trapping and chemical functionalities of additives markedly enhances the performance of LSBs. To the best of our knowledge, this discharge capacity at a high C-rate, e.g. 700 mAh g-1 at 5C, is among the highest in reported data available.50-52

ACS Paragon Plus Environment

14

Page 15 of 23

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 Energy Letters

In summary, we discovered that the heterogeneous catalysis to promote the fragmentation of LiPS based on the ultra-strong adsorption on TiN can markedly improve the discharge capacity and rate capability of lithium sulfur batteries. The TiN- added cathode exhibits 5.5 times higher discharge capacity compared to the conventional cathode even without pretreatment for encapsulating the active materials; and furthermore, when it combines with the carbon interlayer, a discharge capacity reaches to 700 mAh g-1 at a scan rate of 5C, hich is one of the best rate capability results in the present literature. The distinguished catalytic properties of TiN from other Ti-based compounds were revealed to be the ultra-strong chemical bonding between S and the TiN surfaces using the combined computational and experimental analyses. Hence, the TiN surfaces stabilized the short-chain lithium polysulfides preferentially and assists the fragmentation reactions of long-chain polysulfides into shorter chains required and known to be responsible for the slow kinetics of the second discharge plateau. We also confirmed that the heterogenous catalysis approach for LSB can be readily combined with various conventional LiPS capturing techniques like the insertion of carbon mesh or the encapsulation of elemental sulfur to achieve the acceptable cyclability and rate capability for the future commercialization of lithium sulfur batteries ASSOCIATED CONTENT Supporting Information. Experimental section, SEM and EPMA images, additional electrochemical data, UV-Visible spectroscopy, XPS data, computational details and models AUTHOR INFORMATION Corresponding Authors *E-mail: [email protected] (Y. K)

ACS Paragon Plus Environment

15

ACS Energy Letters

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 16 of 23

*E-mail: [email protected] (Y. J) Author Contributions †

T.-G. J, D. S. C and H. S contributed equally to this work.

Notes The authors declare no competing financial interest. ACKNOWLEDGMENT This work was supported by Samsung future technology foundation (SRFC-TA 1403-04)

REFERENCES

(1) 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. (2) Wang, J.; Chew, S. Y.; Zhao, Z. W.; Ashraf, S.; Wexler, D.; Chen, J.; Ng, S. H.; Chou, S. L.; Liu, H. K. Sulfur-mesoporous carbon composites in conjunction with a novel ionic liquid electrolyte for lithium rechargeable batteries. Carbon, 2008, 46, 229-235. (3) Ji, X.; Nazar, L. F. Advances in Li-S batteries. Mater. Chem., 2010, 20, 9821-9826. (4) Su, Y. S.; Manthiram, A. Lithium-sulphur batteries with a microporous carbon paper as a bifunctional interlayer. Nat. commun., 2012, 3, 1166 (5) Mikhaylik, Y. V.; Akridge, J. R. Polysulfide shuttle study in the Li/S battery system. J.

ACS Paragon Plus Environment

16

Page 17 of 23

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 Energy Letters

Electrochem. Soc., 2004, 151, A1969-A1976. (6) Cheon, S.-E.; Ko, K.-S.; Cho, J.-H.; Kim, S.-W.; Chin, E.-Y.; Kim, H.-T. Rechargeable lithium sulfur battery - I. Structural change of sulfur cathode during discharge and charge. J. Electrochem. Soc., 2003, 150, A796-A799. (7) Cheon, S.-E.; Ko, K.-S.; Cho, J.-H.; Kim, S.-W.; Chin, E.-Y.; Kim, H.-T. Rechargeable lithium sulfur battery II. Rate capability and cycle characteristics. J. Electrochem. Soc., 2003, 150, A800-A805. (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) Ji, X.; Lee, K. T.; Nazar, L. F. A highly ordered nanostructured carbon-sulphur cathode for lithium-sulphur batteries. Nat. Mater., 2009, 8, 500-506. (10) Zhang, B.; Qin, X.; Li, G. R.; Gao, X. P. Enhancement of long stability of sulfur cathode by encapsulating sulfur into micropores of carbon spheres. Energ. Environ. Sci., 2010, 3, 1531-1537. (11) 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 LithiumSulfur Batteries. Chem. Int. Edit., 2012, 51, 3591-3595. (12) Jayaprakash, N.; Shen, J.; Moganty, S. S.; Corona, A.; Archer, L. A. Porous Hollow Carbon@Sulfur Composites for High-Power Lithium-Sulfur Batteries. Angew. Chem.

ACS Paragon Plus Environment

17

ACS Energy Letters

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 18 of 23

Int. Ed., 2011, 50, 5904-5908. (13) Song, H.; Yun, S.-W.; Chun, H.-H.; Kim, M.-G.; Chung, K. Y.; Kim, H. S.; Cho, B.W.; Kim, Y.-T. Anomalous Decrease in Structural Disorder Due to Charge Redistribution in Cr-doped Li4Ti5O12 Negative-electrode Materials for High-rate Li-ion Batteries. Energ. Environ. Sci., 2012, 5, 9903-9913. (14) Su, Y.-S.; Fu, Y.; Cochell, T.; Manthiram, A. A strategic approach to recharging lithium-sulphur batteries for long cycle life. Nat. Commun., 2013, 4, 2985. (15) Jeong, T.-G.; Moon, Y. H.; Chun, H.-H.; Kim, H. S.; Cho, B. W.; Kim, Y.-T. Free standing acetylene black mesh to capture dissolved polysulfide in lithium sulfur batteries. Chem. Comm., 2013, 49, 11107-11109. (16) Ji, L.; Rao, M.; Aloni, S.; Wang, L.; Cairns, E. J.; Zhang, Y. Porous carbon nanofibersulfur composite electrodes for lithium/sulfur cells. Energ. Environ. Sci., 2011, 4, 50535059. (17) Zheng, G.; Yang, Y.; Cha, J. J.; Hong, S. S.; Cui, Y. Hollow Carbon NanofiberEncapsulated Sulfur Cathodes for High Specific Capacity Rechargeable Lithium Batteries. Nano Lett., 2011, 11, 4462-4467. (18) Guo, J.; Xu, Y.; Wang, C. Sulfur-Impregnated Disordered Carbon Nanotubes Cathode for Lithium–Sulfur Batteries. Nano Lett., 2011, 11, 4288-4294. (19) Lin, T.; Tang, Y.; Wang, Y.; Bi, H.; Liu, Z.; Huang, F.; Xie, X.; Jiang, M. Scotch-tapelike exfoliation of graphite assisted with elemental sulfur and graphene–sulfur

ACS Paragon Plus Environment

18

Page 19 of 23

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 Energy Letters

composites for high-performance lithium-sulfur batteries. Energ. Environ. Sci., 2013, 6, 1283-1290. (20) Xu, J.; Shui, J.; Wang, J.; Wang, M.; Liu, H.-K.; Dou, S. X.; Jeon, I.-Y.; Seo, J.-M.; Baek, J.-B.; Dai, L. Sulfur–graphene nanostructured cathodes via ball-milling for highperformance lithium–sulfur batteries . ACS nano, 2014, 8, 10920-10930. (21) Moon, S.; Jung, Y. H.; Kim, D. K. Enhanced electrochemical performance of a crosslinked polyaniline-coated graphene oxide-sulfur composite for rechargeable lithium-sulfur batteries. J. Power Sources, 2015, 294, 386-392. (22) Hao, Z.; Yuan, L.; Chen, C.; Xiang, J.; Li, Y.; Huang, Z.; Hu, P.; Huang, Y. TiN as a simple and efficient polysulfide immobilizer for lithium-sulfur batteries. J. Mater. Chem. A, 2016, 4, 17711-17717 (23) Ryu, H. S.; Guo, Z. P.; Ahn, H. J.; Cho, G. B.; Liu, H. K. Investigation of discharge reaction mechanism of lithium vertical bar liquid electrolyte vertical bar sulfur battery. J. Power Sources, 2009, 189, 1179-1183. (24) Demir-Cakan, R.; Morcrette, M.; Nouar, F.; Davoisne, C.; Devic, T.; Gonbeau, D.; Dominko, R.; Serre, C.; Férey, G.; Tarascon, J.-M. Cathode Composites for Li–S Batteries via the Use of Oxygenated Porous Architectures. J. Am. Chem. Soc., 2011, 133, 16154-16160. (25) Ji, X.; Evers, S.; Black, R.; Nazar, L. F. Stabilizing lithium–sulphur cathodes using polysulphide reservoirs. Nat. Commun., 2011, 2, 325.

ACS Paragon Plus Environment

19

ACS Energy Letters

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 20 of 23

(26) Yang, Y.; Yu, G.; Cha, J. J.; Wu, H.; Vosgueritchian, M.; Yao, Y.; Bao, Z.; Cui, Y. Improving the Performance of Lithium–Sulfur Batteries by Conductive Polymer Coating. ACS Nano, 2011, 5, 9187-9193. (27) Tao, X. Y.; Wang, J. G.; Ying, Z. G.; Cai, Q. X.; Zheng, G. Y.; Gan, Y. P.; Huang, H.; Xia, Y.; Liang, C.; Zhang, W. K., et al. Strong Sulfur Binding with Conducting MagneliPhase TinO2n-1 Nanomaterials for Improving Lithium-Sulfur Batteries. Nano Lett., 2014, 14, 5288-5294. (28) Evers, S.; Yim, T.; Nazar, L. F. Understanding the Nature of Absorption/Adsorption in Nanoporous Polysulfide Sorbents for the Li–S Battery. J. Phys. Chem. C, 2012, 116, 19653-19658. (29) Seh, Z. W.; Yu, J. H.; Li, W.; Hsu, P.-C.; Wang, H.; Sun, Y.; Yao, H.; Zhang, Q.; Cui, Y. Two-dimensional layered transition metal disulphides for effective encapsulation of high-capacity lithium sulphide cathodes. Nat. Commun., 2014, 5, 5017. (30) Mosavati, N.; Chitturi, V. R.; Salley, S. O.; Ng, K. Y. S. Nanostructured titanium nitride as a novel cathode for high performance lithium/dissolved polysulfide batteries. J. Power Sources, 2016, 321, 87-93. (31) Salem, H. A.; Babu, G.; Rao, C. V.; Arava, L. M. R. Electrocatalytic polysulfide traps for controlling redox shuttle process of Li-S batteries. J. Am. Chem. Soc., 2015, 137, 11542-11545. (32) Yuan, Z.; Peng, H.-J.; Hou, T.-Z.; Huang, J.-Q.; Chen, C.-M.; Wang, D.-W.; Cheng, X.-B.; Wei, F.; Zhang, Q. Powering lithium-sulfur battery performance by propelling

ACS Paragon Plus Environment

20

Page 21 of 23

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 Energy Letters

polysulfide redox at sulfiphillic hosts. Nano Lett., 2016, 16, 519-527. (33) Balach, J.; Jaumann, T.; Muhlenhoff, S.; Eckert, J.; Giebeler, L. Enhanced polysulphide redox reaction using a RuO2 nanoparticle decorated mesoporous carbon as functional separator coating for advanced lithium-sulphur batteries. Chem. Commun., 2016, 52, 8134-8137. (34) Peng, H. J.; Zhang, Z. W.; Huang, J. Q.; Zhang, G. A cooperative interface for highly efficient lithium-sulfur batteries. Adv. Mater., 2016, 28, 9551-9558. (35) Lu, Y.-C.; He, Q.; Gasteiger, H. A. Probing the Lithium–Sulfur Redox Reactions: A Rotating-Ring Disk Electrode Study. J. Phys. Chem. C, 2014, 118, 5733-5741. (36) Seh, Z. W.; 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. (37) Wild, M.; O'Neill, L.; Zhang, T.; Purkayastha, R.; Minton, G.; Marinescu, M.; Offer, G. J. Lithium sulfur batteries, a mechanistic review. Energ. Environ. Sci., 2015, 8, 34773494. (38) Xu, R.; Lu, J.; Amine, K. Progress in Mechanistic Understanding and Characterization Techniques of Li‐S Batteries. Adv. Energy Mater., 2015, 5, 1500408-1500430. (39) Barchasz, C.; Leprêtre, J.-C.; Alloin, F.; Patoux, S. New insights into the limiting parameters of the Li/S rechargeable cell. J. Power Sources, 2012, 199, 322-330. (40) Cuisinier, M.; Cabelguen, P.-E.; Evers, S.; He, G.; Kolbeck, M.; Garsuch, A.; Bolin, T.;

ACS Paragon Plus Environment

21

ACS Energy Letters

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 22 of 23

Balasubramanian, M.; Nazar, L. F. Sulfur Speciation in Li–S Batteries Determined by Operando X-ray Absorption Spectroscopy. J. Phys. Chem. Lett., 2013, 4, 3227-3232. (41) 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, 3973-3980. (42) Diao, Y.; Xie, K.; Xiong, S.; Hong, X. Shuttle phenomenon–The irreversible oxidation mechanism of sulfur active material in Li–S battery. J. Power Sources, 2013, 235, 181-186. (43) Xiong, S.; Xie, K.; Diao, Y.; Hong, X. Characterization of the solid electrolyte interphase on lithium anode for preventing the shuttle mechanism in lithium–sulfur batteries. J. Power Sources, 2014, 246, 840-845. (44) Zhang, Q.; Wang, Y.; Seh, Z. W.; Fu, Z.; Zhang, R.; Cui, Y. Understanding the Anchoring Effect of Two-Dimensional Layered Materials for Lithium–Sulfur Batteries. Nano Lett., 2015, 15, 3780-3786. (45) Naumkin, A. V.; Kraut-Vass, A.; Gaarenstroom, S. W.; Powell, C. J.; NIST X-ray Photoelectron Spectroscopy Database, Version 4.1; National Institute of Standards and Technology, Gaithersburg,; 2012. (46) Wang, B.; Alhassan, S. M.; Pantelides, S. T. Formation of large polysulfide complexes during the lithium-sulfur battery discharge. Phys. Rev. Appl., 2014, 2, 034004.

ACS Paragon Plus Environment

22

Page 23 of 23

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 Energy Letters

(47) Oyama, S. T. The chemistry of transition metal carbides and nitrides, Springer, 1996. (48) Bartholomew, R. F.; Frankl, D. R. Electrical Properties of Some Titanium Oxides. Physical Review, 1969, 187, 828-833. (49) Vračar, L. M.; Krstajić, N. V.; Radmilović, V. R.; Jakšić, M. M. Electrocatalysis by nanoparticles – oxygen reduction on Ebonex/Pt electrode. J. Electroanal. Chem., 2006, 587, 99-107. (50) Zhao, Q.; Hu, X.; Zhang, K.; Zhang, N.; Hu, Y.; Chen, J. Sulfur Nanodots Electrodeposited on Ni Foam as High-Performance Cathode for Li–S Batteries. Nano Lett., 2015, 15, 721-726. (51) Qiu, Y.; Li, W.; Zhao, W.; Li, G.; Hou, Y.; Liu, M.; Zhou, L.; Ye, F.; Li, H.; Wei, Z. et al. High-Rate, Ultralong Cycle-Life Lithium/Sulfur Batteries Enabled by NitrogenDoped Graphene. Nano Lett., 2014, 14, 4821-4827. (52) Huang, J.-Q.; Zhang, Q.; Zhang, S.-M.; Liu, X.-F.; Zhu, W.; Qian, W.-Z.; Wei, F. Aligned sulfur-coated carbon nanotubes with a polyethylene glycol barrier at one end for use as a high efficiency sulfur cathode. Carbon, 2013, 58, 99-106.

ACS Paragon Plus Environment

23