Metal–Organic Framework as a Multifunctional ... - ACS Publications

May 25, 2018 - The 2032 coin-type cells were fabricated using the. NCM811 electrode as ... planes (2 −1 0) and (3 0 0), respectively. Scanning electro...
3 downloads 0 Views 9MB Size
Research Article pubs.acs.org/journal/ascecg

Cite This: ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Metal−Organic Framework as a Multifunctional Additive for Selectively Trapping Transition-Metal Components in Lithium-Ion Batteries Bum-Jin Chae,†,‡,∥ Yoo Eil Jung,§,∥ Chang Yeon Lee,*,§ and Taeeun Yim*,†,‡ †

Department of Chemistry, Incheon National University, 119, Academy-ro, Yeonsu-gu, Incheon 22012, Republic of Korea Research Institute of Basic Sciences, College of Natural Science, Incheon National University, 119, Academy-ro, Yeonsu-gu, Incheon 22012, Republic of Korea § Department of Energy and Chemical Engineering, Incheon National University, 119, Academy-ro, Yeonsu-gu, Incheon 22012, Republic of Korea ‡

S Supporting Information *

ABSTRACT: To improve the interfacial stability of lithium-ion batteries, a metal−organic framework (MOF) was designed and synthesized as an advanced additive for nickel-rich cathodes to trap the transition metal components. Use of the MOF was found to not compromise the specific capacity of the cells, and cells cycled with a nickel-rich layered oxide embedded with a metal−organic framework exhibited considerably improved cycle retention, even at high temperatures. A systematic analysis demonstrated that only negligible amounts of nickel-ion species migrated from the nickel-rich cathode to the anode surface, and the volume of nickel ions trapped inside the porous structure of the MOF could be determined by quantifying the mass change of the electrode. Finally, the surface degradation triggered by the nickel-ion dissolution was seen to be remarkably suppressed because the MOF improved the surface stability of the nickel-rich cathodes. KEYWORDS: Metal−organic framework, Lithium-ion batteries, Cathode, Additive, Transition metal



during end of cell charging.12−14 However, a Ni-rich NCM cathode is unlikely to exhibit good long-term cycling performance because many unstable Ni4+ species are present on its surface. During the electrochemical process, the electrolyte decomposition in a cell is accelerated by the presence of unstable Ni4+ species, which generates nucleophilic fluoride (F−) species.15−18 The unstable and electrophilic TM components are easily attacked by the nucleophilic F− species via chemical reactions on the Ni-rich NCM cathode surface, leading to the significant dissolution of TM components into the electrolyte.19,20 Note that this reaction is irreversible and permanently deforms the reversible redox reaction sites in the Ni-rich NCM cathode structure. In addition, once the TM components dissolve, they diffuse to the anode side where they promptly precipitate on the anode surface via electrochemical reduction. This reaction disrupts the formation of solid electrolyte interphase (SEI) layers, which are responsible for the stable cycling of the anode.21−23 This means that the TM dissolution not only accelerates the structural deformation of the cathode but also impedes the stable cycling of the anode. Thus, it is important to minimize the loss of TM components

INTRODUCTION With the ever-increasing demand for large-scale electric equipment, such as electric vehicles and energy storage systems, the challenge of achieving a high energy density in lithium-ion batteries (LIBs) has emerged as the most pressing issue currently facing equipment designers.1−6 Since the energy density of a battery cell is significantly affected by the working potential and specific capacity of the electrode material, many studies have investigated advanced electrode materials in the hopes of satisfying the requirements of these new applications. In this regard, a Ni-rich layered oxide (LiNixCoyMnzO2, Ni-rich NCM) has received considerable attention as an alternative cathode material because its specific capacity (>180 mA h g−1) is higher than that of conventional lithium cobalt oxide (150 mA h g−1).7−11 A Ni-rich NCM cathode is composed of (1) lithium ions (Li + ) as charge carriers, (2) transition metal (TM) components, such as Ni, Co, and Mn, to control the electrochemical behavior of the Ni-rich NCM cathode, such as the specific capacity or cycling performance, and (3) oxygen to stabilize the structure of the Ni-rich NCM cathode. In general, increasing the Ni content in this structure can greatly increase its specific capacity because Ni2+, which partially occupies the Co3+ sites, has a lower electrochemical potential than Co3+ and thus affords a much higher specific capacity © XXXX American Chemical Society

Received: February 22, 2018 Revised: May 25, 2018

A

DOI: 10.1021/acssuschemeng.8b00867 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

model in the linear range, as determined using the consistency criteria.32,33 Evaluating the Electrochemical Performance and Analyzing the Cycled Electrodes. The electrode was prepared as follows: First, LiNi0.8Co0.1Mn0.1O2 (Ecopro, NCM811), carbon black (Super P), and PVDF (poly(vinylidene fluoride), KF1100, Kureha) were mixed at a ratio of 90:5:5 in N-methyl-2-pyrrolidone (NMP, Aldrich), and the 2, 5, and 10 wt % MOF was additionally added into the slurry solutions. Each resulting slurry was then separately coated onto thin aluminum foil and dried at 150 °C for 12 h in a vacuum oven. The loading density of the cathode electrode was controlled to be approximately 8.00 mg cm−2. The 2032 coin-type cells were fabricated using the NCM811 electrode as the cathode, lithium metal as the anode, poly(ethylene) (Celgard) as the separator, and an electrolyte (EC:EMC = 1:2 in 1 M LiPF6, PanaxEtec.). The cells were charged to 4.3 V (vs Li/Li+) and discharged to 3.0 V (vs Li/Li+) at 0.1 C for 2 cycles (formation step) and then at 1.0 C for 100 cycles at 25 or 60 °C (Wonatech, WBCS3000). After completing the evaluation, the cycled cells were disassembled in a glovebox. The recovered NCM811 electrodes and anode were washed with dimethyl carbonate (DMC). The thermal behaviors of the cycled NCM811 electrodes were examined by thermogravimetric analysis (TGA, TGA N-1000), and the surface morphology was characterized by transmission electron microscopy (TEM, JEOL) and field-emission scanning electron microscopy (FESEM, JEOL). To quantify the loss of TM components during electrochemical cycling, beaker cells were assembled with the NCM811 cathode and a lithium metal anode, and they were charged to 4.3 V (vs Li/Li+). Then, they were kept in a 60 °C oven for 2 weeks, and the resulting supernatants were analyzed by inductively coupled plasma−mass spectroscopy (ICP−MS, Bruker).

from the cathode structure to enable the fabrication of highperformance cells employing Ni-rich NCM cathodes. To inhibit the dissolution of the TM components in a cell, we propose a metal−organic framework (MOF) as an electrode additive (Figure 1). Recently, MOFs constructed via coordina-

Figure 1. Material strategy employed by the MOF to trap transition metal components.

tion bonds between metal or metal-cluster secondary building units (SBU) and organic building blocks have emerged as new porous materials. These have shown promise in various applications, including gas storage,24 catalysis,25 sensing,26 separation,27 drug delivery,28 and electrocatalysis,29 due to their regularity, enormous surface area, fine-tunability, and confined pore size. In addition, MOFs have been employed as adsorbent materials for the sequestration of metal ions from aqueous solutions,30 and various MOFs have been employed to effectively capture diverse heavy TMs. Strong adsorption or ion exchange sites in the confined pores of MOFs play a pivotal role in the selective sequestration of specific metal ions from a solution. Therefore, MOFs can be employed as an alternative and effective means to prevent TM components from dissolving and diffusing from the cathode to the anode during electrochemical processes, thus stabilizing the surface properties of both electrodes. In this study, we designed a task-specific MOF structure that selectively traps Ni ions based on their size and binding affinity with organic ligands. Then, its electrochemical performance was studied and subsequent systematic analyses were performed to demonstrate the effect of the MOF on the interfacial stability of LIBs. To the best of our knowledge, this is the first attempt to use a task-specific MOF material as an electrode additive to enhance the interfacial stability of a Ni-rich cathode material. Using the MOF as an electrode additive is an attractive approach to improve the interfacial stability of LIBs because the convenient and scalable one-step synthesis process of the MOF allows cells to exhibit a high energy density together with remarkable long-term cycling performance.





RESULTS AND DISCUSSION The Co-MOF-74 was prepared following the procedure published in the literature. The powder X-ray diffraction (PXRD) patterns of the as-synthesized MOF matched well with the simulated patterns, confirming that phase-pure MOF was prepared with the desired characteristics (Figure 2a). The PXRD patterns of the Co-MOF-74 exhibited intense reflections at (2θ angles of) 6.7° and 11.7°, which corresponded to the planes (2 −1 0) and (3 0 0), respectively. Scanning electron microscopy (SEM) images of the Co-MOF-74 are shown in Figure S1 of the Supporting Information (SI). As shown, a growth of large crystals can be seen with a broad distribution of particle sizes and lengths that ranged from 15 to 70 μm in length and 15 to 30 μm in width. A nitrogen (N2) adsorption− desorption isotherm was employed to verify the microporous nature of Co-MOF-74 by highlighting its Type I nature. The Brunauer−Emmett−Teller (BET) surface area of the degassed Co-MOF-74 was measured to be 1122 m2 g−1, which was in excellent agreement with the reported value (Figure 2b). Intriguingly, the pore size distribution of the Co-MOF-74 obtained from HK method exhibited one peak in the 0.4−0.9 nm range with a maximum at 0.51 nm (Figure 2c). This small pore size of the Co-MOF-74 may provide a suitable barrier to suppress the leakage of nickel ions from the cathode. The effects of the MOF additive on the electrochemical performance of the cathodes at room temperature were evaluated, and the results are shown in Figure 3a, b. The initial specific capacities of the cells were almost identical at ∼185.0 mA h g−1 with a Coulombic efficiency of 92.5%. The specific capacity was well maintained after 50 cycles, and all cells retained more than 95.0% of their capacity, indicating that the use of the MOF additive did not compromise the roomtemperature cycling performance of the Ni-rich NCM cathodes. At high temperatures (60 °C), different electrochemical behaviors were observed (Figure 3c, d). During the initial

EXPERIMENTAL SECTION

MOF Synthesis. Co-MOF-74 was synthesized according to a published procedure.31 N2 adsorption/desorption isotherm was measured volumetrically at 77 K in the range of 7.0 × 10−6 ≤ P/P0 ≤ 1.00 with an Autosorb-iQ outfitted with the micropore option from Quantachrome Instruments (Boynton Beach, Florida U.S.A.) with the Autosorb-iQ Win software package. After solvent exchange of the assynthesized materials with MeOH (2 × 10 mL, 12 h each time), the samples were activated (i.e., degassed) at 250 °C for 5 h using the outgas port of the Autosorb-iQ instrument. The specific surface areas for N2 were calculated using the Brunauer−Emmett−Teller (BET) B

DOI: 10.1021/acssuschemeng.8b00867 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

high temperature environment for a longer period of time. These results indicate that the MOF amount is a critical factor in determining the electrochemical performance of the NCM811 cathode, and 5 wt % MOF is the optimum amount for retaining specific capacity. To demonstrate the effectiveness of the MOF additive, the cycled electrodes were subjected to a thermogravimetric analysis (TGA) (Figure S2). The TGA curves of all electrodes exhibited two distinct slopes: one corresponding to the thermal decomposition of the polymeric binder (PVDF), and the other associated with the thermal decomposition of the remaining conducting agent (carbon) on the electrode. In other words, only the inorganic active species of NCM811 were present at the end of the TGA experiments. The mass of the NCM811 additive was 90.0% in the pristine electrode (without MOF) before cycling, and decreased to 83.9% after cycling, which indicated that 6.1% of the TM components were lost during electrochemical cycling. In the electrode with 5 wt % MOF, the mass of the NCM811 additive was 85.7% before cycling and decreased to 83.6% after cycling. The degree of the loss of TM components was 2.1%, which was lower than that obtained for the standard electrode. The ex-situ transmission electron microscopy (TEM) analyses supported these results (Figure 4). In the TEM image of the cycled NCM811 electrode without the MOF, the surface of the NCM811 appeared to be increasingly degraded as the number of cycles increased. Because the dissolution of TM components generally occurs at the surface of the cathode due to the unstable nature at the interface between the electrode and electrolyte,34−36 the surface morphologies can be significantly affected by cycling once irreversible dissolution of the TM components begins. However, the cycled NCM811 with 5 wt % MOF exhibited a well-ordered NCM811 cathode structure. Once the TM components were dissolved in the electrolyte, the surface roughness drastically increased compared with its initial state because the reaction is highly irreversible.34−36 Similar analysis results were obtained from the HAADF images at low magnification for each cycled NCM811 cathode (Figure S3). We confirmed that morphology of primary NCM particles in the cycled 5% MOF-NCM811 was uniform when compared to that before cycling with a constant element ratio of Ni, Co, and Mn (8:1:1). alternatively, the surface state of the primary particles in the cycled NCM811 was not uniform in contrast to the cycled 5% MOF-NCM811, which indicates the primary particles in the cycled NCM811 were increasingly deformed as the number of cycles increased. Note that the cycled NCM811 cathodes with 2% and 10% MOF additive also exhibited relatively clean and smooth surface states compared to that of the cycled cathodes without MOF additive. Less than 10 nm of degraded layers were observed in the MOF-controlled cycled NCM811 cathodes after 50 cycles (Figure S4). This implies that using the MOF as an electrode additive effectively stabilizes the NCM811 structure by selectively trapping the unstable TM components remaining on the NCM811 surface. To quantify the volume of dissolved TM components with and without the MOF, we analyzed the cycled electrolytes using inductively coupled plasma-mass spectroscopy (ICP− MS) (Figure 5). To verify the effectiveness of the MOF, the cells were charged to 4.3 V (vs Li/Li+) and stored at 60 °C for 2 weeks to accelerate the dissolution of the TM components. After this period, the dissolution of the Co and Mn in the cells stored at 60 °C was found to be negligible: 0.92 ppm of Mn and 0.03 ppm of Co were detected in the electrolyte of the cell

Figure 2. (a) XRD patterns of synthesized MOF (red) and simulated MOF (black). (b) N2 adsorption−desorption isotherms for MOF at 77 K, and (c) H−K pore-size distributions of MOF.

cycle, the specific capacity was almost identical regardless of the MOF addition; however, the cycling behavior of the cells depended on how much MOF additive was added. The cycling performance of the cell cycled without the MOF faded continuously, and only 65.5% of the specific capacity retained after 100 cycles. In contrast, the cell cycled with 5 wt % MOF exhibited an improved capacity retention of 82.2% after 100 cycles. This indicates the MOF additive effectively improved the electrochemical performance of the NCM811 cathode. Interestingly, the cell cycled with 2 wt % MOF exhibited relatively poor capacity retention: after 50 cycles, 77.8% of the specific capacity was retained, which is lower than that of the cell cycled without the MOF (83.0%). The cell composed of a 2% MOF-NCM811 cathode also showed stable cycling behavior up to 60 cycles, but the behavior began to fade after 60 cycles. It is surmised that the higher loading of the MOF in the NCM811 cathode increased the scavenging ability for the TM components in the cell, thereby improving the cycling retention. In addition, the pores in the 2% NCM811 cathode appeared to become saturated at 60 cycles. At this point, the MOF no longer trapped the TM components and the electrolyte decomposition process was significantly accelerated because the cells infused with MOF additives were exposed to a C

DOI: 10.1021/acssuschemeng.8b00867 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 3. (a) Charge/discharge potential profiles of NCM811 cathodes at 25 °C and (b) its cycling performances. (c) Charge/discharge potential profiles of NCM811 cathodes at 60 °C and (d) its cycling performances (black: bare NCM811, red: 2% MOF-NCM811, blue: 5% MOF-NCM811, and orange: 10% MOF-NCM811) (bold line: at 1 cycle, and thin line: at 100 cycles).

that did not include MOF, while only 0.54 ppm of Mn and 0.01 ppm of Co were observed in the cell with MOF. This may be attributed to the relatively low contents of Co and Mn in the NCM811 structure. In contrast, the amounts of dissolved Ni were considerably different in the two cases. In the cell fabricated without MOF, a considerable amount of Ni was found to have dissolved in the electrolyte (258.2 ppm), while the cell with 5 wt % MOF exhibited almost 15-times less Ni (18.0 ppm) dissolved in the electrolyte. These results indicate that the MOF effectively scavenges unstable TM components in the cell, thus improving the surface stability and long-term cycling performance of the NCM811 electrode. To demonstrate the effects of the MOF on preventing TM dissolution in the cell, we assembled a full cell comprising NCM811/graphite electrodes and evaluated its electrochemical performance (Figure 6a). Note that once TM dissolution from the surface of the NCM811 cathode began, the dissolved TM were able to easily migrate to the graphite anode and eventually precipitated on the surface of the graphite by electrochemical reduction. When that happened, the surface precipitated TM species disturbed the formation of a uniform SEI layer on the graphite surface (although such a layer is required to maintain the long term cycling performance of the cell), resulting in the rapid fading of the cycling performance. In this respect, this evaluation of a full cell provided sufficient information to estimate the effectiveness of the MOF in the cell and demonstrated that the addition of MOF effectively prevented TM dissolution behavior from the NCM811 cathode. In practice, when the cell with the MOF additive was cycled, the cell retained 78.2% of its capacity after 100 cycles. In contrast,

that without the MOF only retained 58.4% of its capacity. The results of the ICP−MS analyses of the cycled graphite also support this observation (Figure 6b). After 100 cycles, the graphite cycled with the MOF exhibited a relatively low Ni content (174.8 ppm), while the graphite cycled without the MOF showed a large TM content (1138.8 ppm). The surface morphology of the cycled graphite revealed that the use of the MOF effectively stabilized the surface properties of the graphite (Figure 6c, d), and the graphite cycled with the MOF exhibited a relatively clean, uniform surface after one cycle while an irregular surface morphology was observed on the graphite cycled without the MOF. In addition, the surface state of the graphite cycled with the MOF seemed well maintained after 100 cycles, while the surface of the graphite cycled without the MOF was covered with decomposed adducts due to electrolyte decomposition. This result implies that using the MOF effectively stabilizes the surface properties of not only the NCM811 cathode but also the graphite anode by effectively scavenging TM components during the electrochemical process. These results indicate that using the MOF as an electrode additive effectively reduces the concentration of dissolved TM components, and thereby improves the performance of LIBs.



CONCLUSIONS To improve the interfacial stability of Ni-rich cathode material, a task-specific MOF structure was designed and synthesized via a simple one-step process. The MOF additive does not compromise the electrochemical performance of the NCM811 cathode, which retains a remarkable 88.2% of its capacity at D

DOI: 10.1021/acssuschemeng.8b00867 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 4. TEM images of (a) noncycled NCM811, (b) noncycled 5% MOF-NCM811, (c) cycled NCM811, and (d) cycled 5% MOF- NCM811.

Figure 5. Quantification results for dissolved Ni component in the electrolyte during storage test performed at 60 °C for 2 weeks.

high temperatures after 100 cycles. The results of ex-situ TGA analyses of the cycled electrode indicate that the MOF additive effectively suppresses the dissolution of TM components from the NCM811 surface, and additional ICP−MS results indicate that the proposed MOF efficiently traps Ni components in the

cell. These results demonstrate that the proposed MOF structure provides outstanding performance for suppressing Ni dissolution in the cell. We believe that the adopted strategy of designing a task-specific MOF with outstanding electrochemical performance provides insights into expanding the E

DOI: 10.1021/acssuschemeng.8b00867 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering

Figure 6. (a) Cycle performance of NCM811/graphite (black) and 5% MOF-NCM811/graphite (blue), (b) amount of dissolved Ni species observed on surface of graphite anode after 100th charge/discharge, and SEM image of cycled graphite (c) without MOF and (d) with MOF.



ACKNOWLEDGMENTS This work was supported by National Research Foundation of Korea grant from the Korean government (MSIP) (NRF2016R1C1B1009452 and 2017R1A6A1A06015181).

applications of MOF materials, not only in LIBs but also in post-LIB systems that continue to suffer from the deterioration of the electrode material during the electrochemical process.



ASSOCIATED CONTENT



S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.8b00867. SEM and TEM images for MOF are included in Figure S1 and TGA analysis results are presented in Figure S2. TEM images for cycled NCM811 cathodes are shown in Figures S3 and S4 (PDF)



(1) Tarascon, J.-M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359−367. (2) Liu, W.; Oh, P.; Liu, X.; Lee, M.-J.; Cho, W.; Chae, S.; Kim, Y.; Cho, J. Nickel-Rich Layered Lithium Transition-Metal Oxide for HighEnergy Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2015, 54, 4440− 4457. (3) Liu, J.; Zhang, J.-G.; Yang, Z.; Lemmon, J. P.; Imhoff, C.; Graff, G. L.; Li, L.; Hu, J.; Wang, C.; Xiao, J.; Xia, G.; Viswanathan, V. V.; Baskaran, S.; Sprenkle, V.; Li, X.; Shao, Y.; Schwenzer, B. Materials Science and Materials Chemistry for Large Scale Electrochemical Energy Storage: From Transportation to Electrical Grid. Adv. Funct. Mater. 2013, 23, 929−946. (4) Zhang, S.; Wang, M.; Zhou, Z.; Tang, Y. Multifunctional Electrode Design Consisting of 3D Porous Separator Modulated with Patterned Anode for High-Performance Dual-Ion Batteries. Adv. Funct. Mater. 2017, 27, 1703035. (5) Jiang, C.; Fang, Y.; Lang, J.; Tang, Y. Integrated Configuration Design for Ultrafast Rechargeable Dual-Ion Battery. Adv. Energy Mater. 2017, 7, 1700913. (6) Qin, P.; Wang, M.; Li, N.; Zhu, H.; Ding, X.; Tang, Y. BubbleSheet-Like Interface Design with an Ultrastable Solid Electrolyte Layer

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (C.Y.L.). *E-mail: [email protected] (T.Y.). ORCID

Chang Yeon Lee: 0000-0002-1131-9071 Taeeun Yim: 0000-0002-7057-9308 Author Contributions ∥

REFERENCES

These authors contributed equally to this work.

Notes

The authors declare no competing financial interest. F

DOI: 10.1021/acssuschemeng.8b00867 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Research Article

ACS Sustainable Chemistry & Engineering for High-Performance Dual-Ion Batteries. Adv. Mater. 2017, 29, 1606805. (7) Konarov, A.; Myung, S.-T.; Sun, Y.-K. Cathode Materials for Future Electric Vehicles and Energy Storage Systems. ACS Energy Lett. 2017, 2, 703−708. (8) Lee, S.-W.; Kim, M.-S.; Jeong, J. H.; Kim, D.-H.; Chung, K. Y.; Roh, K. C.; Kim, K.-B. Li3PO4 surface coating on Ni-rich LiNi0.6Co0.2Mn0.2O2 by a citric acid assisted sol-gel method: Improved thermal stability and high-voltage performance. J. Power Sources 2017, 360, 206−214. (9) Li, S.; Fu, X.; Zhou, J.; Han, Y.; Qi, P.; Gao, X.; Feng, X.; Wang, B. An effective approach to improve the electrochemical performance of LiNi0.6Co0.2Mn0.2O2 cathode by an MOF-derived coating. J. Mater. Chem. A 2016, 4, 5823−5827. (10) Venkatraman, S.; Choi, J.; Manthiram, A. Factors influencing the chemical lithium extraction rate from layered LiNi1‑y‑zCoyMnzO2 cathodes. Electrochem. Commun. 2004, 6, 832−837. (11) Hua, W.; Zhang, J.; Zheng, Z.; Liu, W.; Peng, X.; Guo, X.-D.; Zhong, B.; Wang, Y.-J.; Wang, X. Na-doped Ni-rich LiNi0.5Co0.2Mn0.3O2 cathode material with both high rate capability and high tap density for lithium ion batteries. Dalton Trans. 2014, 43, 14824−14832. (12) Qiao, R.; Liu, J.; Kourtakis, K.; Roelofs, M. G.; Peterson, D. L.; Duff, J. P.; Deibler, D. T.; Wray, L. A.; Yang, W. Transition-metal redox evolution in LiNi0.5Mn0.3Co0.2O2 electrodes at high potentials. J. Power Sources 2017, 360, 294−300. (13) Kim, J.-M.; Chung, H.-T. The first cycle characteristics of Li[Ni1/3Co1/3Mn1/3]O2 charged up to 4.7 V. Electrochim. Acta 2004, 49, 937−944. (14) Cherkashinin, G.; Motzko, M.; Schulz, N.; Spath, T.; Jaegermann, W. Electron Spectroscopy Study of Li[Ni,Co,Mn]O2/ Electrolyte Interface: Electronic Structure, Interface Composition, and Device Implications. Chem. Mater. 2015, 27, 2875−2887. (15) Liu, W.; Wang, M.; Gao, X. l.; Zhang, W.; Chen, J.; Zhou, H.; Zhang, X. Improvement of the high-temperature, high-voltage cycling performance of LiNi0.5Co0.2Mn0.3O2 cathode with TiO2 coating. J. Alloys Compd. 2012, 543, 181−188. (16) Lee, S.-H.; Yoon, C. S.; Amine, K.; Sun, Y.-K. Improvement of long-term cycling performance of Li[Ni0.8Co0.15Al0.05]O2 by AlF3 coating. J. Power Sources 2013, 234, 201−207. (17) Lee, D.-J.; Scrosati, B.; Sun, Y.-K. Ni3(PO4)2-coated Li[Ni0.8Co0.15Al0.05]O2 lithium battery electrode with improved cycling performance at 55 ◦C. J. Power Sources 2011, 196, 7742−7746. (18) Gallus, D. R.; Schmitz, R.; Wagner, R.; Hoffmann, B.; Nowak, S.; Cekic-Laskovic, I.; Schmitz, R. W.; Winter, M. The influence of different conducting salts on the metal dissolution and capacity fading of NCM cathode material. Electrochim. Acta 2014, 134, 393−398. (19) Jo, C.-H.; Cho, D.-H.; Noh, H.-J.; Yashiro, H.; Sun, Y.-K.; Myung, S. T. An effective method to reduce residual lithium compounds on Ni-rich Li[Ni0.6Co0.2Mn0.2]O2 active material using a phosphoric acid derived Li3PO4 nanolayer. Nano Res. 2015, 8, 1464− 1479. (20) Wang, L.; Mu, D.; Wu, B.; Yang, G.; Gai, L.; Liu, Q.; Fan, Y.; Peng, Y.; Wu, F. Enhanced electrochemical performance of Lithium Metasilicate-coated LiNi0.6Co0.2Mn0.2O2 Ni-rich cathode for Li-ion batteries at high cutoff Voltage. Electrochim. Acta 2016, 222, 806−813. (21) Zheng, H.; Qu, Q.; Zhu, G.; Liu, G.; Battaglia, V. S.; Zheng, H. Quantitative Characterization of the Surface Evolution for LiNi0.5Co0.2Mn0.3O2/Graphite Cell during Long-Term Cycling. ACS Appl. Mater. Interfaces 2017, 9, 12445−12452. (22) He, M.; Su, C.-C.; Feng, Z.; Zeng, L.; Wu, T.; Bedzyk, M. J.; Fenter, P.; Wang, Y.; Zhang, Z. High Voltage LiNi0.5Mn0.3Co0.2O2/ Graphite Cell Cycled at 4.6 V with a FEC/HFDEC-Based Electrolyte. Adv. Energy Mater. 2017, 7, 1700109. (23) Wang, C.; Yu, L.; Fan, W.; Liu, J.; Ouyang, L.; Yang, L.; Zhu, M. 3,3′-(Ethylenedioxy)dipropiononitrile as an Electrolyte Additive for 4.5 V LiNi1/3Co1/3Mn1/3O2/Graphite Cells. ACS Appl. Mater. Interfaces 2017, 9, 9630−9639.

(24) Li, J.-R.; Kuppler, R. J.; Zhou, H.-Ca Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (25) Lee, J. Y.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450−1459. (26) Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Metal-organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105−1125. (27) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal−Organic Frameworks for Separations. Chem. Rev. 2012, 112, 869−932. (28) Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J. F.; Heurtaux, D.; Clayette, P.; Kreuz, C.; Chang, J.-S.; Hwang, Y. K.; Marsaud, V.; Bories, P.-N.; Cynober, L.; Gil, S.; Ferey, G.; Couvreur, P.; Gref, R. Porous metal-organic-framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 2010, 9, 172−178. (29) Morozan, A.; Jaouen, F. Metal organic frameworks for electrochemical applications. Energy Environ. Sci. 2012, 5, 9269−9290. (30) Kobielska, P. A.; Howarth, A. J.; Farha, O. K.; Nayak, S. Metal− organic frameworks for heavy metal removal from water. Coord. Chem. Rev. 2018, 358, 92−107. (31) Caskey, S. R.; Wong-Foy, A. G.; Matzger, A. J. Dramatic Tuning of Carbon Dioxide Uptake via Metal Substitution in a Coordination Polymer with Cylindrical Pores. J. Am. Chem. Soc. 2008, 130, 10870− 10871. (32) Walton, K. S.; Snurr, R. Q. Applicability of the BET Method for Determining Surface Areas of Microporous Metal−Organic Frameworks. J. Am. Chem. Soc. 2007, 129, 8552−8556. (33) Bae, Y.-S.; Yazaydin, A. O.; Snurr, R. Q. Evaluation of the BET Method for Determining Surface Areas of MOFs and Zeolites that Contain Ultra-Micropores. Langmuir 2010, 26, 5475−5483. (34) Banerjee, A.; Ziv, B.; Shilina, Y.; Luski, S.; Aurbach, D.; Halalay, I. C. Acid-Scavenging Separators: A Novel Route for Improving Li-Ion Batteries’ Durability. ACS Energy Lett. 2017, 2, 2388−2393. (35) Hu, P.; Zhao, J.; Wang, T.; Shang, C.; Zhang, J.; Qin, B.; Liu, Z.; Xiong, J.; Cui, G. A composite gel polymer electrolyte with high voltage cyclability for Ni-rich cathode of lithium-ion battery. Electrochem. Commun. 2015, 61, 32−35. (36) Amine, K.; Chen, Z.; Zhang, Z.; Liu, J.; Lu, W.; Qin, Y.; Lu, J.; Curtis, L.; Sun, Y.-K. Mechanism of capacity fade of MCMB/ Li1.1[Ni1/3Mn1/3Co1/3]0.9O2 cell at elevated temperature and additives to improve its cycle life. J. Mater. Chem. 2011, 21, 17754−17759.

G

DOI: 10.1021/acssuschemeng.8b00867 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX