Enhancing the Lithium Storage Capacities of Coordination

Aug 20, 2018 - Key Laboratory of Advanced Energy Materials Chemistry (MOE), College of Chemistry, Nankai University, Tianjin 300071 , China...
6 downloads 0 Views 12MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

Enhancing the Lithium Storage Capacities of Coordination Compounds for Advanced Lithium-Ion Battery Anodes via a Coordination Chemistry Approach Hongwen Liu,† Huanhuan Li,‡ Fangyi Cheng,*,† Wei Shi,*,† Jun Chen,†,‡ and Peng Cheng†,‡,§ †

Department of Chemistry, College of Chemistry, Nankai University, Tianjin 300071, China Key Laboratory of Advanced Energy Materials Chemistry (MOE), College of Chemistry, Nankai University, Tianjin 300071, China § Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Nankai University, Tianjin 300071, China Downloaded via UNIV OF NEW ENGLAND on August 21, 2018 at 00:22:41 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: The influence of the water molecule on both the structural dimensionality and the lithium storage capacities of four coordination compounds was studied. Increasing the reaction temperature to remove the terminal water ligand of discrete coordination compounds [M(HNA)2(H2O)4] (H2NA = 5-hydroxynicotinic acid, M = Co for 1 and Ni for 2) led to forming three-dimensional (3D) coordination polymers [M(NA)]n (M = Co for 3 and Ni for 4). When 1−4 were investigated as active anode materials for lithium storage at 100 mA g−1, the relatively low capacities of 455 and 411 mA h g−1 were obtained after 60 cycles with discrete 1 and 2, while that of 3 and 4 showed high capacities of 618 and 610 mA h g−1 after 100 cycles. Detailed mechanism studies by powder X-ray diffraction, X-ray photoelectron spectroscopy, and scanning electron microscopy showed that the structural dimensionality change induced by water molecules can greatly contribute the cyclability and rate performance for coordination compounds as anode material for lithium storage.



INTRODUCTION Lithium-ion batteries (LIBs) have attracted tremendous attention in the portable electronic market.1−4 As a commercial negative electrode material for lithium ion batteries, graphite has been widely applied. However, the theoretical capacity of this material is 372 mA h g−1, which is hard to meet the increasing requirement for fast development of advanced portable devices. Hence, it has become a research hotspot to search for new anode materials with excellent capacities and cycling abilities. For scientists, to design and synthesize new anode materials rationally for LIBs is still a huge challenge.5−8 Coordination compounds are built by metal centers and organic/inorganic ligands through coordination bonds, which have been well studied in many fields, such as in biochemistry, catalytic chemistry, and porous materials.9,10 As a new subfamily of coordination compounds, coordination polymers (CPs), which are crystalline materials with infinite structures built by organic linkers and inorganic nodes via coordination bonds, have the advantages of tunable structures and versatile functionalities on gas storage and separation,11,12 catalysis,13,14 luminescence,15−17 magnetism,18 chiral recognition,19,20 and so on. CPs have been documented for clean energy applications, such as in fuel cells, LIBs, supercapacitors, and solar cells.21−23 Recently, CPs have been used in LIBs, owning to their structural diversity, adjustable redox properties, simple synthetic methods, and low cost.24,25 However, poor stability © XXXX American Chemical Society

in electrolyte, low electrical conductivity, and large capacity loss upon cycling have prevented their applications in electrochemical applications.26−28 Therefore, it is still very challenging to both obtain high-performance CP-based active materials for LIBs and understand the mechanism of lithiation/ delithiation processes for the rational design of excellent LIBs.29−44 In this contribution, the electrochemical properties of four coordination compounds with 5-hydroxynicotinic acid (H2NA) showing different structural dimensionalities were studied for LIBs. The ligand H2NA was adopted owing to its varied coordination modes and multiple sites for lithium storage.21−23 By removing the terminal water ligands in discrete [M(HNA)2(H2O)4] (M = Co for 1 and Ni for 2), three-dimensional (3D) [M(NA)]n (M = Co for 3 and Ni for 4) formed. When 1−4 were studied as active anode materials for LIBs, discrete 1 and 2 showed relatively low capacities of 455 and 411 mA h g−1 after 60 cycles, while enhanced capacities of 618 and 610 mA h g−1 after 100 cycles for 3 and 4 were obtained. The mechanism for the lithium storage processes of those CPs with different electrochemical behaviors were studied in detail. Received: May 11, 2018

A

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Scheme 1. Synthesis Route of 1−4



for C12H6N2NiO6 4: C, 43.29; H, 1.82; N, 8.41. Found: C, 42.74; H, 1.86; N, 8.43.

EXPERIMENTAL SECTION



Synthesis and Characterization. All raw materials were acquired commercially and utilized directly. The four compounds were measured by the single-crystal X-ray diffraction with graphitemonochromatized Mo Kα radiation (λ = 0.71073 Å) on an Agilent Supernova diffractometer. Elemental analyses were tested by Vario EL cube CHN. The compounds were characterized by Powder X-ray diffraction (PXRD) patterns with Cu Kα radiation on a Rigaku Ultima IV diffractometer. Fourier transform infrared (FTIR) were measured on a Bruker ALPHA spectrophotometer. Thermal stabilities were studied by Thermogravimetric analysis (TGA) from 40 to 800 °C under nitrogen on Labsys NETZSCH TG 209 Setaram equipment. X-ray photoelectron spectroscopy (XPS) with Al Kα Xray source was carried out on a Kratos AXIS Ultra DLD spectrometer. The morphology and microstructure were studied by ZEISS MERLIN Compact scanning electron microscopy (SEM). Electrochemical Measurements. Electrochemical characterizations were evaluated in CR2032 assembled in an Ar-filled glovebox, using lithium metal as the counter/reference electrode, Celgard 2400 membrane as the separate. The electrolyte is 1 M LiPF6 in a mixture of ethylene carbonate (EC) and diethyl carbonate (DEC) (1:1 in volume). The galvanostatic cycling performances were collected on a LAND battery testing system at 298 K. A CHI604E electrochemical workstation was applied to perform cyclic voltammetry (CV) analysis. Mixing active materials, conductive materials (ketjen black), and polyvinylidene fluoride (PVDF) at a weight ratio of 70:20:10 in Nmethyl-2-pyrrolidinone (NMP) solution formed the working electrode. The slurry was pasted onto a copper foil and then dried at 80 °C in vacuum atmosphere for 12 h. The active material loading is about 0.8 mg cm−2. Synthesis. Synthesis of 1 and 2. H2NA (0.0278 g, 0.2 mmol) and CoCl2·6H2O (0.0237 g, 0.1 mmol) were added to 6 mL of H2O− CH3CN (v:v = 2:1) solution in a 10 mL reaction kettle. The kettle was heated to 80 °C and kept for 24 h. The mixture was filtered to obtain a pink microcrystalline material of 1. 2 was obtained using a procedure identical to that of 1 but with NiCl2·6H2O used in place of CoCl2·6H2O. Anal. Calcd for C12H16N2CoO10 1: C, 35.36; H, 3.93; N, 6.88. Found: C, 35.51; H, 3.71; N, 6.92. Calcd for C12H16N2NiO10 2: C, 35.38; H, 3.93; N, 6.88. Found: C, 35.61; H,3.75; N, 6.98. Synthesis of 3 and 4. H2NA (0.0278 g, 0.2 mmol) and CoCl2· 6H2O (0.0237 g, 0.1 mmol) were added to 6 mL of H2O−CH3CN (v:v = 2:1) solvent in a 25 mL reaction kettle. The kettle was placed at 130 °C. After 72 h, it was cooled to room temperature and was filtered to gain a pink product, followed by washing with water at room temperature 2−3 times to remove the unreacted reactants. 4 was obtained using a procedure identical to that of 3 but with NiCl2· 6H2O used in place of CoCl2·6H2O. Anal. Calcd for C12H6N2CoO6 3: C, 43.27; H, 1.82; N, 8.41. Found: C, 42.27; H, 1.84; N, 8.27. Calcd

RESULTS AND DISCUSSION A facile method for the preparation of 1−4 is shown in Scheme 1. The reaction of Co(II) or Ni(II) ions with H2NA at 80 °C produced discrete 1 and 2, which have four terminal water ligands on each metal center. In general, high temperature can reduce the number of the small terminal ligands to enhance the possibility of obtaining high-dimensional frameworks without the aid of any auxiliary ligands.45,46 The reaction of the same raw materials at 130 °C generated 3D coordination polymers of 3 and 4. 1 and 2 and 3 and 4 isostructurally crystallized in the triclinic space group P-1 and the monoclinic space group P21/ c, respectively, studied by single-crystal X-ray diffractions (Table S1). Here the structures of 1 and 3 are depicted as representatives. For 1, as shown in Figure S1a, the asymmetric unit consists of half Co(II) ion, one HNA−, and two coordinated H2O molecules. The metal center is sixcoordinated in a CoO4N2 octahedral coordination environment finished by four oxygens from four H2O molecules and two nitrogens from two HNA−. The Co−O bond length range is 2.068(3) Å to 2.121(3) Å, and the Co−N bond is 2.138(3) Å (Table S2). This coordination environment is similar to a Cd(II) complex.47 The 3D supramolecular structure is maintained through O−H···O hydrogen bonds, as shown in Figure S1b-d. For 3, the asymmetric unit is comprised of one Co(II) ion and one NA2−. The metal center is six-coordinated in a CoO4N2 octahedral configuration completed by two carboxylate oxygens, two hydroxyl oxygens, and two nitrogens from the NA2− (Figure S2a). The bond lengths of Co−O range from 2.065(3) Å to 2.188(3) Å, and the Co−N bond is 2.136(3) Å. Each NA2− anion linked three Co(II) ions through two oxygen atoms and one nitrogen atom, forming a 3D supramolecular structure (Figure S2b). The network topology of 3 was studied by the software package TOPOS20. As shown in Figure S2c, if the metal center is considered as a sixconnected node, NA2− is considered as a three-connected node, and the framework is a (3,6)-connected network with {4·62}{42·610·83} point symbol. PXRD patterns of 1−4 are well consistent with the simulated patterns based on the single-crystal diffraction data, indicative of pure phases of the bulk samples (Figure S3). In the FTIR spectra of 1 and 2, a broad peak of 3500 cm−1 B

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. Charge−discharge curves of 1 (a) and 3 (b) at 100 mA g−1, cycling performance and Coulombic efficiency of 1 (c) and 3 (d) at 100 mA g−1, and rate performance of 1 (e) and 3 (f).

The related composition and structural difference encourage us to study how the impact of structural dimensionality can affect the storage performance of lithium ions. We performed galvanostatic cycling measurements to study the lithium-ion storage property with 1−4 as active materials. As shown in Figures 1a,b and S11a,b, they were tested at 100 mA g−1 in 0.01−3 V. The initial capacities were 1528 and 1642 mA h g−1 for 1 and 2, respectively. The capacities decreased rapidly to 544 and 616 mA h g−1 in the second cycle and maintained stabilized capacities of 455 and 411 mA h g−1 in the following cycles. For 3, the discharge and charge capacities of the first cycle were 1934 and 973 mA h g−1. The capacities of the second cycle decreased to 815 and 775 mA h g−1. Interfacial lithium storage and the formation of a solid electrolyte interphase (SEI) layer with the organic electrolyte decomposed led to the capacity loss.48−52 Another reason for the larger irreversible capacity loss for the initial cycle of 1 and 2 than that of 3 and 4 is induced by the reaction of H2O molecules with lithium ions, which is irreversible. A stabilized capacity of 618 mA h g−1 in the following cycles was achieved.

belongs to O−H vibration of coordinated water molecules (Figure S4a). In the FTIR spectra of 3 and 4 (Figure S4b), the water peaks disappeared. According to the TGA under nitrogen atmosphere (Figure S5a), a weight loss of 17.66% for 1 or 17.68% for 2 was observed from 150 to 250 °C, attributing to the liberation of four coordinated water molecules (calc. 17.68% for 1 and 17.69% for 2). TGA curves of 3 and 4 are similar (Figure S5b). From 150 to 250 °C, no clear weight loss was observed until a remarkable decomposition was found at 350 °C, indicating high thermal stabilities of 3 and 4. To further confirm the thermal stabilities of 3 and 4, we conducted varied temperature PXRD between room temperature and 400 °C (Figure S6). The results indicate that the structural integrity can maintain until 320 and 400 °C for 3 and 4, respectively. Moreover, elemental mapping analysis was carried out for 1−4 by SEM. The elemental mapping results for C, O, N, Co, and Ni element are distributed evenly throughout the structures, indicating the high purity of the compounds (Figures S7−S10). C

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. Cyclic voltammograms at a scan rate of 0.1 mV s−1 in 0.01−3 V for 1 (a), 2 (b), 3 (c), and 4 (d).

To gain insight into the reasons of enhanced electrochemical performance of 3 and 4 in comparison with that of 1 and 2, EIS tests of 1−4 were performed at room temperature (Figure S13). In the high frequencies, the diameter of the depressed semicircle represents the charge transfer. Between the electrode-active component and the liquid electrolyte is charge transfer resistance, which would be influenced by the electrode surface area. Obviously, 3 and 4 exhibit a lower charge transfer resistance than that of 1 and 2, indicating a higher interfacial conductivity. Hence, in 3 and 4, charge transfer becomes much faster, and the redox kinetics of the active materials is improved. Improving the electronic conductivity of 3 and 4 can accelerate the transmission of electrons, which greatly elevate the rate capacities.57 Figure 2 displays the CV profiles of the electrodes. The obvious irreversibility in the initial cycle is ascribed to the SEI layers.48−52 In subsequent cycles, the CV curves are almost overlapping with 1−4, portending a stable cyclic performance. A reversible lithiation potential at 0.62 V was found for 1. Delithiation took place at 1.22 V (Figure 2a). These performances also agreed well with the charge−discharge behaviors of the electrodes at a current density of 100 mA g−1. The oxidation peak appeared at 1.13 V, and the reduction peak located at 0.67 V for 2 was noticed from the second cycle obviously, manifesting an analogical redox process with 1 (Figure 2b). For the CV curves of 3, the broad reduction peak appeared at 0.7 V. Two oxidation peaks were presented to 1.20 and 0.80 V (Figure 2c). For 4, the reduction peak appeared at 0.92 V, and two oxidation peaks were reflected in 1.24 and 1.01 V (Figure 2d). These redox behaviors of 3 and 4 are quite different from that of 1 and 2, due to the different framework structures. The differences also indicate that the lithium storage processes are closely related to microenvironments.53 To investigate the redox reaction mechanism, 1−4 were removed from the electrodes to carry out XPS studies after the

For 4, it has the similar electrochemical behavior of 3. The discharge specific capacity was 1585 mA h g−1 in the first cycle, and a charge specific capacity of 801 mA h g−1 was obtained. The capacities in the second cycle decreased to 670 and 648 mA h g−1, and a capacity of 610 mA h g−1 in the following cycles was stabilized. The cycling performance were further examined. 1 and 2 present relatively low capacities of 455 and 411 mA h g−1 after 60 cycles (Figures 1c,d and S11c,d), whereas the dischargespecific capacities of 3 and 4 can still remain at 618 and 610 mA h g−1 after 100 cycles, suggesting excellent electrochemical reversibility and prominent capacity retention. In comparison to 1 and 2, 3 and 4 performed different redox behaviors, even with the same ligands. These results indicate that the lithium storage processes are closely related to microenvironments of the materials and hence exhibit different cycling stabilities.53 Moreover, the initial Coulombic efficiencies are 35.61%, 29.17%, 50.3%, and 50.51% for 1−4. The Coulombic efficiencies increased to 97.9%, 97.3%, 99.4%, and 99.2% from the second cycle for 1−4, respectively, suggesting that the SEI layers are considerably stable without side reactions.54 The rate performance of 1−4 was tested as the current densities vary (Figure 1e,f). The stable capacities of 3 at 50, 100, 200, 500, and 1000 mA g−1 are 723, 636, 543, 421, and 345 mA h g−1, respectively. The results showed that the anode with 3 is significantly superior than that with 1 (508, 425, 367, 278, and 197 mA h g−1). The capacity returns back to 742 mA h g−1 when the current density turns back to 50 mA g−1 with 3, indicating its fast reaction kinetics and robust electrode integrity.55,56 The high capacities of the anode with 3 are 421 and 345 mA h g−1 at 500 and 1000 mA g−1, respectively, which were also worth noting (Figure S12). In addition, the rate performance of the anode with 3 is superior to that of 1. The anodes with 2 and 4 have a similar rate performance as 1 and 3 (Figure S11e,f). D

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. XPS spectra of 1 (a), (c), 2 (b), (d), 3 (e), (g), and 4 (f), (h).

Co(II) ion of 1 and 3 and the Ni(II) ion of 2 and 4. In Figures S14−S17, the energies of 284.80, 286.12, and 288.15 eV in the C 1s spectra for 1 (284.80, 286.26, 289.60 eV for 2, 284.80, 286.21, 288.49 eV for 3, and 284.80, 285.95, 289.58 eV for 4) are attributed to the CC of the pyridine ring, carbon− oxygen bonds, and carboxyl, respectively, in the electrochemical charge process. The carboxyl transforms into the C− C of the enol structure in the electrochemical discharge process.60,61 During the discharge process, the two peaks of O 1s spectra at 529.78 and 531.21 eV for 1 (529.12, 531.40 eV for 2, 529.45, 530.85 eV for 3, and 529.81, 532.09 eV for 4) correspond to the Li−O bond and the lattice oxygen.62 The N 1s peak of the discharge electrode shifts to a relatively low

charge and discharge process. In the charge state for 1 and 2, the peaks appearing at 781.05 and 797.01 eV and 855.58 and 873.10 eV are attributed to Co(II) and Ni(II) ions (Figure 3a,b).58,59 The Co 2p peaks appeared at 780.98 and 797.12 eV, and the Ni 2p peaks appeared at 856.40 and 874.23 eV in the discharge state (Figure 3c,d). Similarly, the peaks were located at 780.86 and 796.81 eV for Co 2p of 3 and 856.05 and 873.55 for Ni 2p of 4 in the charge state (Figure 3e,f). The Co 2p peaks were reflected in 780.67 and 796.31 eV and 855.35 and 872.49 eV of Ni 2p peaks in the discharge state (Figure 3g, h). The negligibly changed binding energy positions of the peaks and the well-resolved satellite peaks during the process of discharge and charge clearly demonstrate the presence of the E

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. SEM images of 1−4 as electrodes before cycling 1 (a), 2 (b), 3 (e), and 4 (f) and after the electrochemical charge−discharge process of 1 (c), 2 (d), 3 (g), and 4 (h).

moieties and hydroxy oxygen atoms can respectively attach one Li, nitrogen atom can utilize one Li for each, and each double bond in the pyridine ring attaches two Li).64 The possible electrochemical process is depicted in Figure S18. The conductive additive ketjen black contributes a negligible capacity (Figure S19). In addition, 1 and 2 exhibit larger irreversible capacity loss for the initial cycle compared to that of 3 and 4 because of their unique structural features. The 3D coordination polymers perform better electrochemical properties because the coordination bonds formed between the metal centers and organic ligands are much stronger than the

binding energy (from 399.95 to 399.39 eV for 1, 400.06 to 399.28 eV for 2, 399.92 to 399.37 eV for 3, and 399.47 to 399.22 eV for 4) because the electron density was increased by accepting electrons and bonding with lithium ions.63 Based on the above analysis, in these anodes, lithium ions are inserted into the organic moiety without the metal ions participating. Considering 1 mol of NA2− anion can accommodate 8 mol of Li on average, and 1 mol of H2O reacted with 1 mol of Li needed, it is irreversible. As a result, 16 mol of Li was inserted in 1 mol of 1−4. The calculated theoretical capacity based on [M(HNA)2(H2O)4] and [M(NA)]n would be 1053 and 1287 mAh g−1 (carboxylate F

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Inorganic Chemistry interaction of hydrogen bonding, which significantly improved the stability in electrolyte.65 In the charge−discharge processes for lithium storage, structural stability is a key factor for applying CPs as electrode materials. To confirm the structural stabilities of CPs, morphological examination, PXRD, and SEM of 1−4 before and after the electrochemical charge−discharge process were characterized. The retained general morphology of 3 and 4 confirms that they were stable during the lithium-ions de/ insertion process (Figure 4). PXRD demonstrates 3 and 4 can maintain electrode structural integrity while 1 and 2 cannot (Figure S20).

ACKNOWLEDGMENTS



REFERENCES

(1) Choi, J. K.; Aurbach, D. Promise and Reality of Post-LithiumIon Batteries with High Energy Densities. Nat. Rev. Mater. 2016, 1, 16013. (2) Muench, S.; Wild, A.; Friebe, C.; Häupler, B.; Janoschka, T.; Schubert, U. S. Polymer-Based Organic Batteries. Chem. Rev. 2016, 116, 9438−9484. (3) Tang, Y. X.; Zhang, Y. Y.; Li, W. L.; Ma, B.; Chen, X. D. Rational material design for ultrafast rechargeable lithium-ion batteries. Chem. Soc. Rev. 2015, 44, 5926−5940. (4) Armand, M.; Tarascon, J. M. Building better batteries. Nature 2008, 451, 652−657. (5) Schon, T. B.; McAllister, B. T.; Li, P. F.; Seferos, D. S. The rise of organic electrode materials for energy storage. Chem. Soc. Rev. 2016, 45, 6345−6404. (6) Xia, W.; Mahmood, A.; Zou, R. Q.; Xu, Q. Metal-organic frameworks and their derived nanostructures for electrochemical energy storage and conversion. Energy Environ. Sci. 2015, 8, 1837− 1866. (7) Zheng, M. B.; Tang, H.; Li, L. L.; Hu, Q.; Zhang, L.; Xue, H. G.; Pang, H. Hierarchically Nanostructured Transition Metal Oxides for Lithium-Ion Batteries. Adv. Sci. 2018, 5, 1700592. (8) Chen, K. S.; Balla, I.; Luu, N. S.; Hersam, M. C. Emerging Opportunities for Two-Dimensional Materials in Lithium-Ion Batteries. ACS Energy Lett. 2017, 2, 2026−2034. (9) Constable, E. C.; Housecroft, C. E. Coordination chemistry: the scientific legacy of Alfred Werner. Chem. Soc. Rev. 2013, 42, 1429− 1439. (10) Werner, H. Alfred Werner: A Forerunner to Modern Inorganic Chemistry. Angew. Chem., Int. Ed. 2013, 52, 6146−6153. (11) Li, J. R.; Kuppler, R. J.; Zhou, H. C. Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (12) Murray, L. J.; Dinca, M.; Long, J. R. Hydrogen storage in metalorganic frameworks. Chem. Soc. Rev. 2009, 38, 1294−1314. (13) Ma, L. Q.; Abney, C.; Lin, W. B. Enantioselective catalysis with homochiral metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1248−1256. (14) Zhu, L.; Liu, X. Q.; Jiang, H. L.; Sun, L. B. Metal-Organic Frameworks for Heterogeneous Basic Catalysis. Chem. Rev. 2017, 117, 8129−8176. (15) Hu, Z. C.; Deibert, B. J.; Li, J. Luminescent metal-organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815−5840. (16) Lustig, W. P.; Mukherjee, S.; Rudd, N. D.; Desai, A. V.; Li, J.; Ghosh, S. K. Metal-organic frameworks: functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242−3285. (17) Cui, Y. J.; Yue, Y. F.; Qian, G. D.; Chen, B. L. Luminescent Functional Metal-Organic Frameworks. Chem. Rev. 2012, 112, 1126− 1162. (18) Liu, K.; Zhang, X. J.; Meng, X. X.; Shi, W.; Cheng, P.; Powell, A. K. Constraining the coordination geometries of lanthanide centers and magnetic building blocks in frameworks: a new strategy for molecular nanomagnets. Chem. Soc. Rev. 2016, 45, 2423−2439. (19) Xue, M.; Li, B.; Qiu, S. L.; Chen, B. L. Emerging functional chiral microporous materials: synthetic strategies and enantioselective separations. Mater. Today 2016, 19, 503−515. (20) Li, B.; Wen, H. M.; Cui, Y. J.; Zhou, W.; Qian, G. D.; Chen, B. L. Emerging Multifunctional Metal-Organic Framework Materials. Adv. Mater. 2016, 28, 8819−8860.

CONCLUSIONS In conclusion, the structural dimensionality changes of four cobalt/nickel-based coordination compounds can greatly influence the electrochemical properties for lithium-ion batteries when they were investigated as active materials. The 3D coordination polymers perform better electrochemical properties than the discrete coordination compounds because the coordination bonds forming the framework between the metal centers and organic ligands are much stronger than the interaction of hydrogen bonding in the discrete system, which improved the stability of the materials in electrolyte during the electrochemical charge−discharge process. The absence of water molecules in the framework compounds also has a positive effect on the improved electrochemical properties. As a result, the stabilized reversible capacity of 618 and 610 mA h g−1 after 100 cycles and the nearly 100% Coulombic efficiency were obtained in the anodes with 3D coordination polymers 3 and 4. We anticipate that coordination compounds with welldesigned metal and organic components can inspire a new direction for high-performance anode materials of LIBs. ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b01295. PXRD, FTIR, TGA, and SEM images and elemental mappings of 1−4, additional electrochemical diagrams and XPS spectra, and EIS measurements (PDF) Accession Codes

CCDC 1830866 and 1830868−1830870 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/ cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.





This work was supported by the National Natural Science Foundation of China (grant numbers 21622105, 21421001, and 21501071) and the Ministry of Education of China (Grant number B12015).





Article

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Fangyi Cheng: 0000-0002-9400-1500 Wei Shi: 0000-0001-6130-1227 Jun Chen: 0000-0001-8604-9689 Peng Cheng: 0000-0003-0396-1846 Notes

The authors declare no competing financial interest. G

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (21) Morozan, A.; Jaouen, F. Metal organic frameworks for electrochemical applications. Energy Environ. Sci. 2012, 5, 9269−9290. (22) Zhou, J. W.; Wang, B. Emerging crystalline porous materials as a multifunctional platform for electrochemical energy storage. Chem. Soc. Rev. 2017, 46, 6927−6945. (23) Wu, H. B.; Lou, X. W. Metal-organic frameworks and their derived materials for electrochemical energy storage and conversion: Promises and challenges. Sci. Adv. 2017, 3, eaap9252. (24) Wang, L.; Han, Y. Z.; Feng, X.; Zhou, J. W.; Qi, P. F.; Wang, B. Metal-organic frameworks for energy storage: Batteries and supercapacitors. Coord. Chem. Rev. 2016, 307, 361−381. (25) Zheng, S. S.; Li, X. R.; Yan, B. Y.; Hu, Q.; Xu, Y. X.; Xiao, X.; Xue, H. G.; Pang, H. Transition-Metal (Fe, Co, Ni) Based MetalOrganic Frameworks for Electrochemical Energy Storage. Adv. Energy Mater. 2017, 7, 1602733. (26) Li, X. X.; Cheng, F. Y.; Zhang, S. N.; Chen, J. Shape-controlled synthesis and lithium-storage study of metal-organic frameworks Zn4O(1,3,5-benzenetribenzoate)2. J. Power Sources 2006, 160, 542− 547. (27) Ferey, G.; Millange, F.; Morcrette, M.; Serre, C.; Doublet, M. L.; Greneche, J. M.; Tarascon, J. M. Mixed-Valence Li/Fe-Based Metal-Organic Frameworks with Both Reversible Redox and Sorption Properties. Angew. Chem., Int. Ed. 2007, 46, 3259−3263. (28) Zhang, Z. Y.; Yoshikawa, H.; Awaga, K. Monitoring the SolidState Electrochemistry of Cu(2,7-AQDC) (AQDC = Anthraquinone Dicarboxylate) in a Lithium Battery: Coexistence of Metal and Ligand Redox Activities in a Metal-Organic Framework. J. Am. Chem. Soc. 2014, 136, 16112−16115. (29) Eftekhari, A. Potassium secondary cell based on Prussian blue cathode. J. Power Sources 2004, 126, 221−228. (30) Armand, M.; Grugeon, S.; Vezin, H.; Laruelle, S.; Ribiere, P.; Poizot, P.; Tarascon, J. M. Conjugated dicarboxylate anodes for Li-ion batteries. Nat. Mater. 2009, 8, 120−125. (31) Morita, Y.; Nishida, S.; Murata, T.; Moriguchi, M.; Ueda, A.; Satoh, M.; Arifuku, K.; Sato, K.; Takui, T. Organic tailored batteries materials using stable open-shell molecules with degenerate frontier orbitals. Nat. Mater. 2011, 10, 947−951. (32) Wessells, C. D.; Peddada, S. V.; Huggins, R. A.; Cui, Y. Nickel Hexacyanoferrate Nanoparticle Electrodes For Aqueous Sodium and Potassium Ion Batteries. Nano Lett. 2011, 11, 5421−5425. (33) Nokami, T.; Matsuo, T.; Inatomi, Y.; Hojo, N.; Tsukagoshi, T.; Yoshizawa, H.; Shimizu, A.; Kuramoto, H.; Komae, K.; Tsuyama, H.; Yoshida, J. I. Polymer-Bound Pyrene-4,5,9,10-tetraone for FastCharge and -Discharge Lithium-Ion Batteries with High Capacity. J. Am. Chem. Soc. 2012, 134, 19694−19700. (34) Zhao, L.; Zhao, J. M.; Hu, Y. S.; Li, H.; Zhou, Z. B.; Armand, M.; Chen, L. Q. Disodium Terephthalate (Na2C8H4O4) as High Performance Anode Material for Low-Cost Room-Temperature Sodium-Ion Battery. Adv. Energy Mater. 2012, 2, 962−965. (35) Liang, Y. L.; Zhang, P.; Chen, J. Function-oriented design of conjugated carbonyl compound electrodes for high energy lithium batteries. Chem. Sci. 2013, 4, 1330−1337. (36) Wang, S. W.; Wang, L. J.; Zhang, K.; Zhu, Z. Q.; Tao, Z. L.; Chen, J. Organic Li4C8H2O6 Nanosheets for Lithium-Ion Batteries. Nano Lett. 2013, 13, 4404−4409. (37) Wang, L.; Lu, Y. H.; Liu, J.; Xu, M. W.; Cheng, J. G.; Zhang, D. W.; Goodenough, J. B. A Superior Low-Cost Cathode for a Na-Ion Battery. Angew. Chem., Int. Ed. 2013, 52, 1964−1967. (38) Wang, R. Y.; Wessells, C. D.; Huggins, R. A.; Cui, Y. Highly Reversible Open Framework Nanoscale Electrodes for Divalent Ion Batteries. Nano Lett. 2013, 13, 5748−5752. (39) You, Y.; Wu, X. L.; Yin, Y. X.; Guo, Y. G. High-quality Prussian blue crystals as superior cathode materials for room-temperature sodium-ion batteries. Energy Environ. Sci. 2014, 7, 1643−1647. (40) Wang, S. W.; Wang, L. J.; Zhu, Z. Q.; Hu, Z.; Zhao, Q.; Chen, J. All Organic Sodium-Ion Batteries with Na4C8H2O. Angew. Chem. 2014, 126, 6002−6006. (41) Li, G. H.; Yang, H.; Li, F. C.; Cheng, F. Y.; Shi, W.; Chen, J.; Cheng, P. A Coordination Chemistry Approach for Lithium-Ion

Batteries: The Coexistence of Metal and Ligand Redox Activities in a One-Dimensional Metal-Organic Material. Inorg. Chem. 2016, 55, 4935−4940. (42) Zhang, L.; Cheng, F. Y.; Shi, W.; Chen, J.; Cheng, P. Transition-Metal-Triggered High-Efficiency Lithium Ion Storage via Coordination Interactions with Redox-Active Croconate in OneDimensional Metal-Organic Anode Materials. ACS Appl. Mater. Interfaces 2018, 10, 6398−6406. (43) Dong, C. F.; Xu, L. Q. Cobalt- and Cadmium-Based MetalOrganic Frameworks as High-Performance Anodes for Sodium Ion Batteries and Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 7160−7168. (44) Hou, Y. P.; Mao, H. Z.; Xu, L. Q. MIL-100(V) and MIL100(V)/rGO with various valence states of vanadium ions as sulfur cathode hosts for lithium-sulfur batteries. Nano Res. 2017, 10, 344− 353. (45) Moulton, B.; Zaworotko, M. J. From Molecules to Crystal Engineering: Supramolecular Isomerismand Polymorphismin Network Solids. Chem. Rev. 2001, 101, 1629−1658. (46) Ockwig, N. W.; Friedrichs, O. D.; O’Keeffe, M.; Yaghi, O. M. Reticular Chemistry: Occurrence and Taxonomy of Nets and Grammar for the Design of Frameworks. Acc. Chem. Res. 2005, 38, 176−182. (47) Jiang, M. X.; Feng, Y. L. Tetraaquabis(5-hydroxynicotinato-KN)-cadmium(II). Acta Crystallogr., Sect. E: Struct. Rep. Online 2008, 64, m1517. (48) Nam, K. T.; Kim, D. W.; Yoo, P. J.; Chiang, C. Y.; Meethong, N.; Hammond, P. T.; Chiang, Y. M.; Belcher, A. M. Virus-Enabled Synthesis and Assembly of Nanowires for Lithium Ion Battery Electrodes. Science 2006, 312, 885−888. (49) 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. (50) Gou, L.; Hao, L. M.; Shi, Y. X.; Ma, S. L.; Fan, X. Y.; Xu, L.; Li, D. L.; Wang, K. One-pot synthesis of a metal-organic framework as an anode for Li-ion batteries with improved capacity and cycling stability. J. Solid State Chem. 2014, 210, 121−124. (51) Song, H. W.; Shen, L. S.; Wang, J.; Wang, C. X. Reversible lithiation-delithiation chemistry in cobalt based metal organic framework nanowire electrode engineering for advanced lithium-ion batteries. J. Mater. Chem. A 2016, 4, 15411−15419. (52) Chen, X. X.; Wang, Z.; Zhang, R. R.; Xu, L. Q.; Sun, D. A novel polyoxometalate-based hybrid containing a 2D [CoMo 8O26]∞ structure as the anode for lithium-ion batteries. Chem. Commun. 2017, 53, 10560−10563. (53) Gong, T.; Lou, X. B.; Gao, E. Q.; Hu, B. W. Pillared-Layer Metal-Organic Frameworks for Improved Lithium-Ion Storage Performance. ACS Appl. Mater. Interfaces 2017, 9, 21839−21847. (54) Wang, Y.; Qu, Q. T.; Liu, G.; Battaglia, V. S.; Zheng, H. H. Aluminum fumarate-based metal organic frameworks with tremellalike structure as ultrafast and stable anode for lithium-ion batteries. Nano Energy 2017, 39, 200−210. (55) Zhou, W. W.; Zhu, J. X.; Cheng, C. W.; Liu, J. P.; Yang, H. P.; Cong, C. X.; Guan, C.; Jia, X. T.; Fan, H. J.; Yan, Q. Y. A general strategy toward graphene@metal oxide core-shell nanostructures for high-performance lithium storage. Energy Environ. Sci. 2011, 4, 4954− 4961. (56) He, C. N.; Wu, S.; Zhao, N. Q.; Shi, C. S.; Liu, E. Z.; Li, J. J. Carbon-Encapsulated Fe3O4 Nanoparticles as a High-Rate Lithium Ion Battery Anode Material. ACS Nano 2013, 7, 4459−4469. (57) Wang, Y.; Deng, Y. H.; Qu, Q. T.; Zheng, X. Y.; Zhang, J. Y.; Liu, G.; Battaglia, V. S.; Zheng, H. H. Ultrahigh-Capacity Organic Anode with High-Rate Capability and Long Cycle Life for LithiumIon Batteries. ACS Energy Lett. 2017, 2, 2140−2148. (58) Sun, C. C.; Yang, J.; Rui, X. H.; Zhang, W. N.; Yan, Q. Y.; Chen, P.; Huo, F. W.; Huang, W.; Dong, X. C. MOF-directed templating synthesis of a porous multicomponent dodecahedron with H

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry hollow interiors for enhanced lithium-ion battery anodes. J. Mater. Chem. A 2015, 3, 8483−8488. (59) Grosvenor, A. P.; Biesinger, M. C.; Smart, R. St. C.; McIntyre, N. S. New interpretations of XPS spectra of nickel metal and oxides. Surf. Sci. 2006, 600, 1771−1779. (60) Markevich, E.; Salitra, G.; Aurbach, D. Fluoroethylene Carbonate as an Important Component for the Formation of an Effective Solid Electrolyte Interphase on Anodes and Cathodes for Advanced Li-Ion Batteries. ACS Energy Lett. 2017, 2, 1337−1345. (61) Beamson, G.; Briggs, D. High resolution XPS of organic polymers: the Scienta ESCA300 database; Wiley: Chichester, UK, 1992. (62) Marco, J. F.; Gancedo, J. R.; Gracia, M.; Gautier, J. L.; Ríos, E.; Berry, F. J. Characterization of the nickel cobaltite, NiCo2O4, prepared by several methods: an XRD, XANES, EXAFS, and XPS study. J. Solid State Chem. 2000, 153, 74−81. (63) Tian, B. B.; Ding, Z. J.; Ning, G. H.; Tang, W.; Peng, C. X.; Liu, B.; Su, J.; Su, C. L.; Loh, K. P. Amino group enhanced phenazine derivatives as electrode materials for lithium storage. Chem. Commun. 2017, 53, 2914−2917. (64) Han, X. Y.; Yi, F.; Sun, T. L.; Sun, J. T. Synthesis and electrochemical performance of Li and Ni 1,4,5,8-naphthalenetetracarboxylates as anodes for Li-ion batteries. Electrochem. Commun. 2012, 25, 136−139. (65) Zhang, H. B.; Wu, S. T.; Tian, C. B.; Lin, Z. J.; Li, Z. H.; Lin, P.; Du, S. W. Two enantiomorphic 3D Zn(II)-carboxylate MOFs with double helical structures serving as a chiral source induced by hydrogen bonding. CrystEngComm 2012, 14, 4165−4167.

I

DOI: 10.1021/acs.inorgchem.8b01295 Inorg. Chem. XXXX, XXX, XXX−XXX