Fluorine Doping Strengthens the Lithium-Storage Properties of the Mn

Jul 26, 2017 - Key Laboratory of Eco-Environment-Related Polymer Materials of Ministry of Education, Key Laboratory of Polymer Materials of Gansu Prov...
0 downloads 13 Views 4MB Size
Research Article www.acsami.org

Fluorine Doping Strengthens the Lithium-Storage Properties of the Mn-Based Metal−Organic Framework Shuhua He,†,‡ Xiaozhong Zhou,§ Zhangpeng Li,† Jinqing Wang,*,† Limin Ma,† and Shengrong Yang*,† †

State Key Laboratory of Solid Lubrication, Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou 730000, P. R. China ‡ University of Chinese Academy of Sciences, Beijing 100080, P. R. China § Key Laboratory of Eco-Environment-Related Polymer Materials of Ministry of Education, Key Laboratory of Polymer Materials of Gansu Province, College of Chemistry and Chemical Engineering, Northwest Normal University, Lanzhou 730000, P. R. China S Supporting Information *

ABSTRACT: The electrochemical properties of the metal−organic framework (MOF)-based composite as electrode material can be significantly improved by means of partial destruction of the full coordination of linkers to metal ions and replacing with other small ions, which make metal centers become more accostable and consequently more effective for the lithiation/ delithiation process. In this paper, F− was chosen to replace some of the benzenedicarboxylate (BDC) linkers because of its better interaction with the Li+ than the oxide ion. What’s more, the formed M−F bond promotes the Li+ to transfer at the active material interface and protects the surface from HF attacking. The as-synthesized F-doped Mn-MOF electrode maintains a reversible capacity of 927 mA h g−1 with capacity retention of 78.5% after 100 cycles at 100 mA g−1 and also exhibits a high discharge capacity of 716 mA h g−1 at 300 mA g−1 and 620 mA h g−1 at 500 mA g−1 after 500 cycles. Even at 1000 mA g−1, the electrode still maintains a high reversible capacity of 494 mA h g−1 after 500 cycles as well as a Coulombic efficiency of nearly 100%, which is drastically increased compared with pure Mn-MOF material as expected. KEYWORDS: metal−organic frameworks, fluorine doping, varied fluorine contents, accessible metal sites, lithium storage

1. INTRODUCTION

In particular, MOFs have emerged as novel electrode materials in LIBs because the redox-active metal centers can accommodate multiple electrons24 and the linker structure promotes charge transfer inside the framework.9,25 However, low electrical conductivity26 and poor stability of MOFs in the electrolyte27 have hampered their widespread practical realization for energy storage. Thus, significant research efforts have been concentrated on enhancing the performance of MOFs toward practical applications, such as preparing composites of MOFs with conductive phases,28−30 functionalization and modification of the ligands to construct new MOFs,31,32 introduction of open metal sites,33−35 and so on. It has been demonstrated in a number of studies that introducing

Research and development of high-performance electrode materials for rechargeable lithium-ion batteries (LIBs) are of vital importance to meet the ever-increasing energy storage demands for a wide range of applications.1 Even though considerable research efforts have been devoted to the traditional metal-based inorganic compound electrode materials, redox-active organic electrode materials, like lithium carboxylates,2−4 quinones,5−8 anhydrides,9 organosulfur compounds,10−12 metal−organic frameworks (MOFs),13−17etc., have emerged as a promising alternative due to their recyclability, structural diversity, flexibility, as well as easy functionalization for improving the electrochemical performance.18−21 Among those, various MOFs have attracted considerable attention in recent decades due to their structural diversities and potential applications on various domains.22,23 © 2017 American Chemical Society

Received: June 1, 2017 Accepted: July 26, 2017 Published: July 26, 2017 26907

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914

Research Article

ACS Applied Materials & Interfaces

Figure 1. (a) XRD patterns, (b) Raman spectra, (c) FT-IR spectra, and (d) TG curves of MOF and F-MOF-3 samples. anhydrous methanol to remove residual DMF, followed by filtering and freeze-drying. The obtained samples are named after F-MOF-n (n = 1, 2, ..., 7) based on the amounts of HF. 2.2. Characterization. X-ray power diffraction (XRD, Rigaku D/ Max-2400 diffractometer with Cu Kα radiation), Fourier transform infrared spectroscopy (FT-IR, IFS66 V, Bruker), and Raman spectra (Renishaw 2000 laser Raman spectrometer at 532 nm) were used to investigate the structures of the prepared samples. The chemical composition was investigated by X-ray photoelectron spectroscopy (XPS, EscaLab 250Xi) using an Al Kα X-ray source. The morphology and microstructure of the samples were performed using a scanning electron microscope (SEM, JSM-6701F, JEOL, Japan) and a transmission electron microscope (TEM, JEOL JEM-2010), respectively. Thermal gravimetric analysis (TGA, STA 449C, Germany) of the samples was carried out with a heating rate of 10 °C min−1 from 50 to 800 °C under N2 flow. 2.3. Electrochemical Measurements. To evaluate the electrochemical performances of the prepared F-MOF materials, galvanostatic charge−discharge (GCD), cyclic voltammetry (CV), and electrochemical impedance spectroscopy (EIS) measurements were provided. The GCD experiments were performed in the voltage window of 0.01−3 V vs Li+/Li using a battery test system (LAND CT2001A, China). The EIS measurement was performed using the electrochemical workstation (CHI 660C) from 10 kHz to 0.1 Hz with an alternate current amplitude of 5 mV. The working electrodes were fabricated by spreading a slurry comprising of active material F-MOF, carbon black, and polyvinylidene fluoride (PVDF) with a mass ratio of 7:2:1 dissolved in Nmethylpyrrolidinone (NMP) on the Cu foil current collector. The electrodes were punched into 10 mm diameter discs and then were assembled into coin type cells (CR2032) with Li metal foil as the counter/reference electrode and a polypropylene membrane (Celgard 2400) as the cell separator, while 1 M LiPF6 dissolved in ethylene carbonate (EC)/ethylmethyl carbonate (EMC)/dimethyl carbonate (DMC) (1:1:1 v/v/v) was the electrolyte in the argon-filled glovebox. The cycling performance was first tested at the current density of 100 mA g−1 to select the sample with optimum performance. Figure S1 shows that the sample prepared with 3 mmol HF (denoted as FMOF-3) has the highest reversible capacity. Then, the F-MOF-3 sample’s structure and electrochemical performances were studied in detail. The F-MOF-3 sample was galvanostatically charged and discharged to different cutoff voltages at 20 mA g−1 to gain insight

open metal sites is an effective way to tune the property of MOFs.34,36−38 For example, the catalytic property of the ZrBDC can be improved significantly due to the creation of large quantities of unsaturated Zr sites which are accessible for catalytic reactions.34 Additionally, replacement of some BDC linkers with electron-withdrawing groups can make the resulting MOF materials more active.37 As we all know, fluorine (F) is the most electronegative element, and the F− presents a stronger interaction with the Li+ than the oxide ion; therefore, it is anticipated that the stronger M−F bond at the material surface can promote the Li+ transfer at the active material interface and protect the surface from HF attacking,39 which is beneficial to improving the structural stability of the MOF-based materials. Inspired by these motivational works, Mn-MOF (abbreviated as MOF) materials doped with varied fluorine contents are elaborately designed and prepared under solvothermal conditions aimed at investigating the effects of fluorine doping on the structure and electrochemical performances of MnMOF materials, and there are no correlative reports involved to date as far as we know. Delightfully, the obtained F-doped MnMOF (abbreviated as F-MOF) materials show enhanced thermal stability and conductivity; especially, their electrochemical properties are greatly improved compared with pure Mn-MOF electrode as expected when used as the anode material of LIBs.

2. EXPERIMENTAL SECTION 2.1. Preparation of F-MOF Series. All chemicals are analytical grade and used directly. F-MOF samples were prepared through the solvothermal method. At first, 4 mmol of Mn(CH3COO)2·4H2O was dissolved in 20 mL of N,N′- dimethylformamide (DMF). Next 30 mL of DMF solution containing 4 mmol of 1,4-benzenedicarboxylic acid (H2BDC) was subsequently added into the above solution slowly under constant stirring. Then, various amounts of hydrofluoric acid (HF, Prolabo, 40%, 0−7 mmol) were added. After being stirred for a few minutes, the resulting suspension was introduced into the Teflonlined steel autoclave and heated to 160 °C for 20 h. Finally, the resulting F-MOF precipitates were washed repeatedly with ample 26908

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914

Research Article

ACS Applied Materials & Interfaces

bond,45 which is conducive for improving the electrochemical performance and the safety of rechargeable batteries. The detailed elemental composition information on MOF and F-MOF-3 samples were clearly acquired from the XPS characterization (Figure 2). Both XPS survey spectra possess characteristic peaks of C 1s, O 1s, and Mn 2p, while an additional F 1s signal is also detected in F-MOF-3. In the highresolution spectra of Mn 2p, two major peaks of Mn 2p3/2 and Mn 2p1/2 are located at higher binding energies after fluorine doping, further providing evidence of forming a M−F bond. Figure 3 presents the SEM and TEM images of the prepared MOF and F-MOF-3 samples. Obviously, the morphology of the as-prepared materials has been changed after F doping. Furthermore, the mapping analysis of C, O, F, and Mn elements from the designated area reveals the uniform distribution of these elements within the samples, as clearly shown in Figure 3e. To evaluate the electrochemical properties of the MOF and F-MOF electrodes, the cycling performance was first tested at a moderate current density of 100 mA g−1 from 0.01 to 3.0 V versus Li metal reference/counter electrode. Remarkably, the obtained F-MOF materials all exhibit the enhanced electrochemical performance, and the doping amount of fluorine has a significant influence on the electrochemical performance of FMOF electrodes (Figure S1). Namely, modest fluoride doping into the molecular structure can largely enhance the properties of the material, whereas excess fluorination will form the inert fluoride-based interface films, which can lead to the deterioration of the battery performance.46,47 Specifically, the F-MOF-3 electrode exhibits the highest reversible capacity and maintains a reversible capacity of 927 mA h g−1 with capacity retention of 78.5% after 100 cycles (Figure 4a), which was greatly improved compared with MOF electrode (30 mA h g−1 after 100 cycles). This may be partly put down to the enhanced electrical conductivity of F-MOF-3 material. The EIS measurements in Figure 4b confirm that the F-MOF-3 electrode has a smaller semicircle diameter than that of the MOF electrode in the high frequency, suggesting that the F-MOF-3 electrode possesses lower contact and charge-transfer resistance. Besides, the diffusion resistance is also lower as the slope of the low frequency line increases after fluorine doping.48,49 These results indicate that the electron transport and Li+ diffusion speed during the electrochemical lithiation and delithiation reaction can be enhanced through fluorine doping, which was consistent with previous studies50,51 and thus significantly enhance the electrochemical dynamic behavior.52 Figure 4c shows the charge−discharge curves of the F-MOF3 electrode at a current density of 100 mA g−1. The large irreversible capacity loss arising in the initial cycle mainly results from the solid electrolyte interface (SEI) film formation,4,53 resulting in low Coulombic efficiencies. The corresponding dQ/dV differential curve for the first cycle of F-MOF-3 is shown in Figure 4d. The first cathodic peak at 1.21 V is attributed to the SEI film formation, whereas the latter two cathodic peaks around 0.67 and 0.095 V are ascribed to the reduction of Mn2+ to Mn054 and lithiation of the unsaturated carbons in organic ligands,4 respectively. The two anodic peaks around 0.83 and 1.15 V correspond to the delithiation of the organic ligands and oxidation of Mn0. Under the combined actions of the reduction of the transition metal and subsequent lithiation of the organic ligands, high reversible capacities are obtained, which is consistent with the work reported by Hong and co-workers.4

into the lithiation mechanism of the F-MOF material. The cycled cells were disassembled in the argon-filled glovebox and naturally dried. The electrode slices were used for ex situ XRD, XPS, and Raman tests directly while the powders were scraped from the electrode slices for ex situ FT-IR analysis.

3. RESULTS AND DISCUSSION The XRD patterns of the MOF and F-MOF-3 samples are shown in Figure 1a. Obviously, all the diffraction peaks of MOF can be well assigned to the monoclinic structure (C2/m), which is similar to that of Cu(tpa)·(dmf) reported by Tannenbaum et al.40 After fluorine doping, the crystal structure changes to orthorhombic structure (Cmmm (no. 65)) similar to that of VIII2(OH)2F2{O2C−C6H4−CO2}·H2O synthesized by Férey et al.,41 and the main product is proved to be MnF2 when the HF content is 7 mmol (Figure S2). In Raman analysis (Figure 1b), the peaks appearing at 1610, 1413, and 1135 cm−1 are resulting from the in-plane vibrational modes of the aromatic rings, while the peaks at 858 and 629 cm−1 are due to the out-of-plane deformation of the C−H groups in BDC2− ligands.42−44 The FT-IR analysis (Figure 1c) further verifies the chemical structure variation of MOF after fluorine doping, since it is observed that the band of OH− (3406 cm−1) disappeared while some characteristic peaks of H2BDC and free hydroxyl (3610 cm−1) appeared after fluorine doping, indicating that part of BDC linkers has been replaced by F− successfully. The simplified schematic illustration for forming F-MOF structure is depicted in Scheme 1. Scheme 1. Schematic Illustration for Forming F-MOF Structure

The TG measurement was performed in N2 from 50 to 800 °C to investigate the effect of fluorine doping on the thermal stability of the MOF materials. As can be seen from Figure 1d, the initial weight loss of 16% for MOF at around 160 °C is assigned to the removal of water molecules. In the case of FMOF-3, there is no obvious weight loss below 200 °C, indicating that the water molecules have been substituted by F atoms. Subsequently, the weight losses (39% for MOF and 42% for F-MOF-3) are due to the combustion of the terephthalic acid of the framework. It is important to highlight that the increased thermal stability may be explained by the factor that F− with stronger electronegativity offers a stronger Mn−F 26909

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914

Research Article

ACS Applied Materials & Interfaces

Figure 2. (a) XPS survey spectra and (b) high-resolution Mn 2p spectra of MOF and F-MOF-3 samples.

Figure 3. SEM and TEM images of (a and c) MOF and (b and d) F-MOF-3 samples. (e) EDS elemental mapping image of F-MOF-3 from the selected region.

H3). Besides, the characteristic peaks of vC−H vibrations of phenyls also disappeared after the lithiation process, implying the lithiation of the organic ligands.9 In contrast, a new characteristic absorption band appeared at 862 cm−1 from state H2, which can be attributed to the formation of the SEI film (Li2CO3). Identical conclusions were also obtained in the ex situ Raman analysis where the peaks of benzene ring bending in plane (635 cm−1), benzene ring stretching, and OCO bending of carboxylate in plane (856 cm−1)43 also disappeared, and the intensity of other typical peaks of the benzene ring decreased after the lithiation process. When subsequently charged to 3 V, no obvious peaks can be observed, which is indicative of the

Based on the above assumption, the possible electrochemical process for Li ion insertion may be expressed as shown in Scheme 2.4,9,55 Ex situ XPS, FT-IR, Raman, and XRD analyses for the FMOF-3 electrode at different lithiation/delithiation states (as shown in Figure 5a) were conducted to verify the proposed Li storage mechanism above. As can be seen from the ex situ FTIR spectra of Figure 5b, the peaks at 1595 and 1348 cm−1, resulting from the carboxylate CO stretching and C−O stretching and the sharp absorption peaks at 1499 and 1409 cm−1, which can be attributed to the CC stretching of benzene ring, all disappeared after the lithiation process (state 26910

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914

Research Article

ACS Applied Materials & Interfaces

Figure 4. (a) Cycling performance of the MOF and F-MOF-3 electrodes at current density of 100 mA g−1. (b) Impedance plots of MOF and FMOF-3 electrodes. (c) Potential profiles in cycles 1, 2, 5, 10, 20, and 50 and (d) dQ/dV differential curve for the first cycle of F-MOF-3 electrode.

good stability of the F-MOF-3 electrode. The slight augment of capacity may be ascribed to the reversible growth of the gel-like polymer layer,56,57 coinciding with the cyclic performance results (Figure 4a). Comprehensive cycling performance at various current rates was further evaluated. As shown in Figure 6b, this electrode exhibits a discharge capacity of 716 mA h g−1 at 300 mA g−1 and 620 mA h g−1 at 500 mA g−1 after 500 cycles. Even at 1000 mA g−1, the electrode still retains a high reversible capacity of 494 mA h g−1 after 500 cycles and a Coulombic efficiency of nearly 100%, implying the superiority of the F-MOF-3 electrode in holding electrode stability. Compared with the recent studies on LIBs with MOF-based composites as electrode materials (Table S1), the F-MOF material in this work possesses high capacity and high electrochemical cycling stability. The superior electrochemical properties of F-MOF materials is attributed to the following factors: (i) part of BDC linkers’ missing break the full coordination of the framework metal ions, which makes metal centers become more accostable and thus more effective for electrochemical reactions;34 (ii) F doping can create abundant defects and reactive sites to facilitate the diffusion of Li ion;49,58 (iii) the F− has a stronger interaction with the Li+ than the oxide ion and the M−F bond at the active material surface is expected to promote the Li+ ion diffusion at the active material/electrolyte solution interface; (iv) the M−F bond can protect the active material surface against the attack of HF, which results from the hydrolysis of the electrolyte salt of LiPF6 with a minute amount of water in the cell.39

Scheme 2. Proposed Reaction Mechanism of F-MOF-3 Electrode upon the Electrochemical Lithiation

formation of amorphous phases. This is supported by the observation from the ex situ XRD patterns, as shown in Figure 5d, indicating the initial lithiation/delithiation cycle is a progressive amorphization process, with losing of long-range order of molecular packing.4,9 When the sample is discharged to state H2, the XPS spectrum of Mn 2p does not change obviously compared with that of the pristine F-MOF-3 (H1), demonstrating the chemical state of the manganese element does not change and the formation of the SEI film does not affect the structure of the sample, as shown in Figure S3. After the first discharge (H3), the XPS spectrum of the sample is different from that of the state H1 performed, indicating the chemical state variation of the manganese element. In addition, the Mn 2p spectrum after the charge process (H4) is consistent with that of the original F-MOF-3 sample, suggesting the reversibility of the lithiation/ delithiation process as well as the valence variation of Mn during the charge−discharge process. The rate capacities of the F-MOF-3 electrode were also carried out in the voltage window of 0.01−3.0 V versus Li+/Li. As depicted in Figure 6a, the F-MOF-3 electrode delivers discharge capacities of 736, 794, 848, 764, and 548 mA h g−1 at increasing rates of 100, 200, 300, 500, and 1000 mA g−1, respectively, and resumes a high capacity of about 901 mA h g−1 when returned to 100 mA g−1, which also confirms the

4. CONCLUSIONS Partial destruction of the full coordination of linkers to metal ions and replacing with electron-withdrawing F− make the resulting MOF materials significantly more active. Besides, the greater interaction of F− with the Li+ and the formed M−F bond will promote the Li+ transfer at the active material interface and protect the surface from HF attacking. Furthermore, the enhanced thermal stability and conductivity 26911

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914

Research Article

ACS Applied Materials & Interfaces

Figure 5. Lithiation-delithiation analysis of the F-MOF-3 electrode: (a) a potential-composition profile at 20 mA g−1. Ex situ analyses of FT-IR spectra (b), Raman spectra (c), and XRD pattern (d) at different charge−discharge states. H1, an initial potential (2.2 V); H2, discharged (0.94 V); H3, discharged (0.01 V); and H4, charged (3.0 V) in (a).

Figure 6. (a) Rate performance of the F-MOF-3 electrode at varying rates from 100 to 1000 mA g−1 and (b) cycling performance tested at 300, 500, and 1000 mA g−1, respectively.

of the F-MOF electrode result in drastically increased electrochemical properties. In particular, this F-MOF electrode exhibits high capacity, impressive rate capability, and extremely stable cycle performance. More importantly, our strategy may also be extended to other MOF-based anodes to achieve superior performance for next-generation energy storage and other applications.





sample at different charge−discharge states, and table of MOFs as lithium-ion battery anode. (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. Tel.: +86 931 4968076. Fax: +86 931 8277088. *E-mail: [email protected].

ASSOCIATED CONTENT

S Supporting Information *

ORCID

Xiaozhong Zhou: 0000-0001-5366-0545 Jinqing Wang: 0000-0002-0768-6960

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.7b07843. Cycling performance and XRD patterns of the F-MOF series, ex situ Mn 2p XPS spectra of the F-MOF-3

Notes

The authors declare no competing financial interest. 26912

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914

Research Article

ACS Applied Materials & Interfaces



Dimensional Metal-Organic Material. Inorg. Chem. 2016, 55, 4935− 4940. (18) Liang, Y.; Tao, Z.; Chen, J. Organic Electrode Materials for Rechargeable Lithium Batteries. Adv. Energy Mater. 2012, 2, 742−769. (19) Schon, T. B.; McAllister, B. T.; Li, P.; Seferos, D. S. The Rise of Organic Electrode Materials for Energy Storage. Chem. Soc. Rev. 2016, 45, 6345−6404. (20) Xie, J.; Zhang, Q. Recent Progress in Rechargeable Lithium Batteries with Organic Materials as Promising Electrodes. J. Mater. Chem. A 2016, 4, 7091−7106. (21) Song, Z.; Zhou, H. Towards Sustainable and Versatile Energy Storage Devices: An Overview of Organic Electrode Materials. Energy Environ. Sci. 2013, 6, 2280−2301. (22) Xuan, W.; Zhu, C.; Liu, Y.; Cui, Y. Mesoporous Metal-organic Framework Materials. Chem. Soc. Rev. 2012, 41, 1677−1695. (23) Silva, P.; Vilela, S. M. F.; Tomé, J. P. C.; Paz, F. A. A. Multifunctional Metal-Organic Frameworks: From Academia to Industrial Applications. Chem. Soc. Rev. 2015, 44, 6774−6803. (24) Zhang, Y.; Cheng, T.; Wang, Y.; Lai, W.; Pang, H.; Huang, W. A Simple Approach to Boost Capacitance: Flexible Supercapacitors Based on Manganese Oxides@MOFs via Chemically Induced In Situ Self-Transformation. Adv. Mater. 2016, 28, 5242−5248. (25) Kobayashi, Y.; Jacobs, B.; Allendorf, M. D.; Long, J. R. Conductivity, Doping, and Redox Chemistry of a Microporous Dithiolene-Based Metal-Organic Framework. Chem. Mater. 2010, 22, 4120−4122. (26) Férey, G.; Millange, F.; Morcrette, M.; Serre, C.; Doublet, M. L.; Greneche, J. M.; Tarascon, J. M. Mixed-Valence Li/Fe-Based MetalOrganic Frameworks with Both Reversible Redox and Sorption Properties. Angew. Chem., Int. Ed. 2007, 46, 3259−3263. (27) Kumar, R. S.; Nithya, C.; Gopukumar, S.; Kulandainathan, M. A. Diamondoid-Structured Cu-Dicarboxylate-based Metal-Organic Frameworks as High-Capacity Anodes for Lithium-Ion Storage. Energy Technol. 2014, 2, 921−927. (28) Banerjee, P. C.; Lobo, D. E.; Middag, R.; Ng, W. K.; Shaibani, M. E.; Majumder, M. Electrochemical Capacitance of Ni-doped Metal Organic Framework and Reduced Graphene Oxide Composites: More than the Sum of Its Parts. ACS Appl. Mater. Interfaces 2015, 7, 3655− 3664. (29) Srimuk, P.; Luanwuthi, S.; Krittayavathananon, A.; Sawangphruk, M. Solid-Type Supercapacitor of Reduced Graphene Oxide-Metal Organic Framework Composite Coated on Carbon Fiber Paper. Electrochim. Acta 2015, 157, 69−77. (30) Zhao, Z.; Wang, S.; Liang, R.; Li, Z.; Shi, Z.; Chen, G. Graphene-Wrapped Chromium-MOF(MIL-101)/Sulfur Composite for Performance Improvement of High-Rate Rechargeable Li-S Batteries. J. Mater. Chem. A 2014, 2, 13509−13512. (31) Chang, G.; Wen, H.; Li, B.; Zhou, W.; Wang, H.; Alfooty, K.; Bao, Z.; Chen, B. A Fluorinated Metal-Organic Framework for High Methane Storage at Room Temperature. Cryst. Growth Des. 2016, 16, 3395−3399. (32) Zhang, D.; Chang, Z.; Li, Y.; Jiang, Z.; Xuan, Z.; Zhang, Y.; Li, J.; Chen, Q.; Hu, T.; Bu, X. Fluorous Metal-Organic Frameworks With Enhanced Stability and High H2/CO2 Storage Capacities. Sci. Rep. 2013, 3, 3312. (33) Hong, D. Y.; Hwang, Y. K.; Serre, C.; Férey, G.; Chang, J. S. Porous Chromium Terephthalate MIL-101 with Coordinatively Unsaturated Sites: Surface Functionalization, Encapsulation, Sorption and Catalysis. Adv. Funct. Mater. 2009, 19, 1537−1552. (34) Vandichel, M.; Hajek, J.; Vermoortele, F.; Waroquier, M.; De Vos, D. E.; Van Speybroeck, V. Active Site Engineering in UiO-66 Type Metal-Organic Frameworks by Intentional Creation of Defects: A Theoretical Rationalization. CrystEngComm 2015, 17, 395−406. (35) Wu, H.; Zhou, W.; Yildirim, T. High-Capacity Methane Storage in Metal-Organic Frameworks M2(dhtp): The Important Role of Open Metal Sites. J. Am. Chem. Soc. 2009, 131, 4995−5000. (36) Valenzano, L.; Civalleri, B.; Chavan, S.; Bordiga, S.; Nilsen, M. H.; Jakobsen, S.; Lillerud, K. P.; Lamberti, C. Disclosing the Complex Structure of UiO-66 Metal Organic Framework: A Synergic

ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (No. 51502306).



REFERENCES

(1) Goodenough, J. B.; Kim, Y. Challenges for Rechargeable Li Batteries. Chem. Mater. 2010, 22, 587−603. (2) Armand, M.; Grugeon, S.; Vezin, H.; Laruelle, S.; Ribière, P.; Poizot, P.; Tarascon, J. M. Conjugated Dicarboxylate Anodes for LiIon Batteries. Nat. Mater. 2009, 8, 120−125. (3) Walker, W.; Grugeon, S.; Vezin, H.; Laruelle, S.; Armand, M.; Wudl, F.; Tarascon, J. M. Electrochemical Characterization of Lithium 4,4′-tolane-dicarboxylate for Use as a Negative Electrode in Li-Ion Batteries. J. Mater. Chem. 2011, 21, 1615−1620. (4) Lee, H. H.; Park, Y.; Kim, S. H.; Yeon, S. H.; Kwak, S. K.; Lee, K. T.; Hong, S. Y. Mechanistic Studies of Transition Metal-Terephthalate Coordination Complexes upon Electrochemical Lithiation and Delithiation. Adv. Funct. Mater. 2015, 25, 4859−4866. (5) Song, Z.; Qian, Y.; Gordin, M. L.; Tang, D.; Xu, T.; Otani, M.; Zhan, H.; Zhou, H.; Wang, D. Polyanthraquinone as a Reliable Organic Electrode for Stable and Fast Lithium Storage. Angew. Chem. 2015, 127, 14153−14157. (6) Chen, H.; Armand, M.; Courty, M.; Jiang, M.; Grey, C. P.; Dolhem, F.; Tarascon, J. M.; Poizot, P. Lithium Salt of Tetrahydroxybenzoquinone: Toward the Development of a Sustainable Li-Ion Battery. J. Am. Chem. Soc. 2009, 131, 8984−8988. (7) Song, Z.; Xu, T.; Gordin, M. L.; Jiang, Y.; Bae, I. T.; Xiao, Q.; Zhan, H.; Liu, J.; Wang, D. Polymer-Graphene Nanocomposites as Ultrafast Charge and Discharge Cathodes for Rechargeable Lithium Batteries. Nano Lett. 2012, 12, 2205−2211. (8) Shimizu, A.; Kuramoto, H.; Tsujii, Y.; Nokami, T.; Inatomi, Y.; Hojo, N.; Suzuki, H.; Yoshida, J. I. Introduction of Two Lithiooxycarbonyl Groups Enhances Cyclability of Lithium Batteries with Organic Cathode Materials. J. Power Sources 2014, 260, 211−217. (9) Han, X.; Qing, G.; Sun, J.; Sun, T. How Many Lithium Ions Can be Inserted onto Fused C6 Aromatic Ring Systems? Angew. Chem., Int. Ed. 2012, 51, 5147−5151. (10) Zhang, J.; Kong, L.; Zhan, L.; Tang, J.; Zhan, H.; Zhou, Y.; Zhan, C. Sulfides Organic Polymer: Novel Cathode Active Material for Rechargeable Lithium Batteries. J. Power Sources 2007, 168, 278−281. (11) Zhan, L.; Song, Z.; Shan, N.; Zhang, J.; Tang, J.; Zhan, H.; Zhou, Y.; Li, Z.; Zhan, C. Poly(tetrahydrobenzodithiophene): High Discharge Specific Capacity as Cathode Material for Lithium Batteries. J. Power Sources 2009, 193, 859−863. (12) Burkhardt, S. E.; Conte, S.; Rodriguez Calero, G. G.; Lowe, M. A.; Qian, H.; Zhou, W.; Gao, J.; Hennig, R. G.; Abruña, H. D. Towards Organic Energy Storage: Characterization of 2,5-bis(methylthio)thieno[3,2-b]thiophene. J. Mater. Chem. 2011, 21, 9553−9563. (13) Song, H.; Shen, L.; Wang, J.; Wang, C. Reversible LithiationDelithiation Chemistry in Cobalt Based Metal Organic Framework Nanowire Electrode Engineering for Advanced Lithium-Ion Batteries. J. Mater. Chem. A 2016, 4, 15411−15419. (14) Li, C.; Lou, X.; Shen, M.; Hu, X.; Guo, Z.; Wang, Y.; Hu, B.; Chen, Q. High Anodic Performance of Co 1,3,5-Benzenetricarboxylate Coordination Polymers for Li-Ion Battery. ACS Appl. Mater. Interfaces 2016, 8, 15352−15360. (15) Maiti, S.; Pramanik, A.; Manju, U.; Mahanty, S. Cu3(1,3,5benzenetricarboxylate)2 Metal-Organic Framework: A promising Anode Material for Lithium-Ion Battery. Microporous Mesoporous Mater. 2016, 226, 353−359. (16) Maiti, S.; Pramanik, A.; Manju, U.; Mahanty, S. Reversible Lithium Storage in Manganese 1,3,5-Benzenetricarboxylate MetalOrganic Framework with High Capacity and Rate Performance. ACS Appl. Mater. Interfaces 2015, 7, 16357−16363. (17) Li, G.; Yang, H.; Li, F.; Cheng, F.; Shi, W.; Chen, J.; Cheng, P. A Coordination Chemistry Approach for Lithium-Ion Batteries: The Coexistence of Metal and Ligand Redox Activities in a One26913

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914

Research Article

ACS Applied Materials & Interfaces Combination of Experiment and Theory. Chem. Mater. 2011, 23, 1700−1718. (37) Vermoortele, F.; Vandichel, M.; Van de Voorde, B.; Ameloot, R.; Waroquier, M.; Van Speybroeck, V.; De Vos, D. E. Electronic Effects of Linker Substitution on Lewis Acid Catalysis with MetalOrganic Frameworks. Angew. Chem., Int. Ed. 2012, 51, 4887−4890. (38) Vermoortele, F.; Ameloot, R.; Vimont, A.; Serre, C.; De Vos, D. An Amino-Modified Zr-terephthalate Metal-Organic Framework as an Acid-Base Catalyst for Cross-Aldol Condensation. Chem. Commun. 2011, 47, 1521−1523. (39) Nakajima, T.; Groult, H. Fluorinated Materials for Energy Conversion; Elsevier: London, 2005. (40) Carson, C. G.; Hardcastle, K.; Schwartz, J.; Liu, X.; Hoffmann, C.; Gerhardt, R. A.; Tannenbaum, R. Synthesis and Structure Characterization of Copper Terephthalate Metal-Organic Frameworks. Eur. J. Inorg. Chem. 2009, 2009, 2338−2343. (41) Barthelet, K.; Adil, K.; Millange, F.; Serre, C.; Riou, D.; Férey, G. Synthesis, Structure Determination and Magnetic Behaviour of the First Porous Hybrid Oxyfluorinated Vanado(III)carboxylate: MIL-71 or VIII2(OH)2F2{O2C-C6H4-CO2}·H2O. J. Mater. Chem. 2003, 13, 2208−2212. (42) Bordiga, S.; Lamberti, C.; Ricchiardi, G.; Regli, L.; Bonion, F.; Damin, A.; Lillerud, K. P.; Bjorgen, M.; Zecchina, A.; Bonino, F. Electronic and Vibrational Properties of a MOF-5 Metal-Organic Framework: ZnO Quantum Dot Behaviour. Chem. Commun. 2004, 2300−2301. (43) Hu, Y.; Zhang, L. Amorphization of Metal-Organic Framework MOF-5 at Unusually Low Applied Pressure. Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81, 174103. (44) He, S.; Li, Z.; Wang, J.; Wen, P.; Gao, J.; Ma, L.; Yang, Z.; Yang, S. MOF-Derived NixCo1−x(OH)2 Composite Microspheres for HighPerformance Supercapacitors. RSC Adv. 2016, 6, 49478−49486. (45) Zhang, Z.; Xu, L.; Jiao, H. Ionothermal Synthesis, Structures, Properties of Cobalt-1,4-benzenedicarboxylate Metal-Organic Frameworks. J. Solid State Chem. 2016, 238, 217−222. (46) Yonezawa, S.; Yamasaki, M.; Takashima, M. Surface Fluorination of the Cathode Active Materials for Lithium Secondary Battery. J. Fluorine Chem. 2004, 125, 1657−1661. (47) Amatucci, G. G.; Pereira, N. Fluoride Based Electrode Materials for Advanced Energy Storage Devices. J. Fluorine Chem. 2007, 128, 243−262. (48) Dubal, D. P.; Gund, G. S.; Lokhande, C. D.; Holze, R. Decoration of Spongelike Ni(OH)2 Nanoparticles onto MWCNTs Using an Easily Manipulated Chemical Protocol for Supercapacitors. ACS Appl. Mater. Interfaces 2013, 5, 2446−2454. (49) Zhan, L.; Yang, S.; Wang, Y.; Wang, Y.; Ling, L.; Feng, X. Fabrication of Fully Fluorinated Graphene Nanosheets Towards HighPerformance Lithium Storage. Adv. Mater. Interfaces 2014, 1, 1300149. (50) Boegeat, D.; Jousseaume, B.; Toupance, T.; Campet, G.; Fournès, L. The First Mixed-Valence Fluorotin Alkoxides: New SolGel Precursors of Fluorine-Doped Tin Oxide Materials. Inorg. Chem. 2000, 39, 3924−3927. (51) Kwon, C. W.; Campet, G.; Portier, J.; Poquet, A.; Fournes, L.; Labrugere, C.; Jousseaume, B.; Toupance, T.; Choy, J. H.; Subramanian, M. A. A New Single Molecular Precursor Route to Fluorine-Doped Nanocrystalline Tin Oxide Anodes for Lithium Batteries. Int. J. Inorg. Mater. 2001, 3, 211−214. (52) Jiang, X.; Yu, W.; Wang, H.; Xu, H.; Liu, X.; Ding, Y. Enhancing the Performance of MnO by Double Carbon Modification for Advanced Lithium-Ion Battery Anodes. J. Mater. Chem. A 2016, 4, 920−925. (53) Lin, Y.; Zhang, Q.; Zhao, C.; Li, H.; Kong, C.; Shen, C.; Chen, L. An Exceptionally Stable Functionalized Metal-Organic Framework for Lithium Storage. Chem. Commun. 2015, 51, 697−699. (54) Rui, K.; Wen, Z. Y.; Lu, Y.; Jin, J.; Shen, C. One-Step Solvothermal Synthesis of Nanostructured Manganese Fluoride as an Anode for Rechargeable Lithium-Ion Batteries and Insights into the Conversion Mechanism. Adv. Energy Mater. 2015, 5, 1401716.

(55) Combelles, C.; Yahia, M. B.; Pedesseau, L.; Doublet, M. L. Design of Electrode Materials for Lithium-Ion Batteries: The Example of Metal-Organic Frameworks. J. Phys. Chem. C 2010, 114, 9518− 9527. (56) Li, C.; Hu, X.; Lou, X.; Chen, Q.; Hu, B. Bimetallic Coordination Polymer as a Promising Anode Material for LithiumIon Batteries. Chem. Commun. 2016, 52, 2035−2038. (57) Cai, Z.; Xu, L.; Yan, M.; Han, C.; He, L.; Hercule, K. M.; Niu, C.; Yuan, Z.; Xu, W.; Qu, L.; Zhao, K.; Mai, L. Manganese Oxide/ Carbon Yolk-Shell Nanorod Anodes for High Capacity Lithium Batteries. Nano Lett. 2015, 15, 738−744. (58) Adcock, J. L.; Fulvio, P. F.; Dai, S. Towards the Selective Modification of Soft-Templated Mesoporous Carbon Materials by Elemental Fluorine for Energy Storage Devices. J. Mater. Chem. A 2013, 1, 9327−9331.

26914

DOI: 10.1021/acsami.7b07843 ACS Appl. Mater. Interfaces 2017, 9, 26907−26914