Interconnected Ni(HCO3)2 Hollow Spheres Enabled by Self-Sacrificial

Dec 9, 2016 - (3, 4) Transition-metal oxides (MxOy; M = Fe, Co, Mn, and Ni) have been ..... S. J.; Paik , U.; Lou , X. Construction of Hybrid Bowl-Lik...
0 downloads 0 Views 1MB Size
Subscriber access provided by Olson Library | Northern Michigan University

Letter

Interconnected Ni(HCO3)2 Hollow Spheres Enabled by SelfSacrificial Templating with Enhanced Lithium Storage Properties Shiqiang Zhao, Zewei Wang, Yanjie He, Beibei Jiang, Yeu Wei Harn, Xueqin Liu, Faqi Yu, Fan Feng, Qiang Shen, and Zhiqun Lin ACS Energy Lett., Just Accepted Manuscript • DOI: 10.1021/acsenergylett.6b00582 • Publication Date (Web): 09 Dec 2016 Downloaded from http://pubs.acs.org on December 10, 2016

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

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

Page 1 of 13

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

ACS Energy Letters

Interconnected Ni(HCO3)2 Hollow Spheres Enabled by Self-Sacrificial Templating with Enhanced Lithium Storage Properties Shiqiang Zhao,†‡ Zewei Wang,‡ Yanjie He,‡ Beibei Jiang,‡ Yeuwei Harn,‡ Xueqin Liu,‡ Faqi Yu,† Fan Feng,† Qiang Shen,*† and Zhiqun Lin*‡ †

Key Laboratory for Colloid and Interface Chemistry of Education Ministry, School of Chemistry and

Chemical Engineering, Shandong University, Jinan 250100, PR China ‡

School of Materials Science and Engineering, Georgia Institute of Technology, Atlanta, GA 30332,

USA Corresponding Author * E-mail: [email protected] (Q. Shen). * E-mail: [email protected] (Z. Lin)

ACS Paragon Plus Environment

1

ACS Energy Letters

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

Page 2 of 13

ABSTRACT: Interconnected nickel bicarbonate (Ni(HCO3)2) hollow spheres were produced and exploited for the first time as an anode of lithium ion batteries, delivering the 80th reversible capacity of 1442 mAh g-1 at the current rate of 100 mA g-1 which is 3.9 times the theoretical capacity of commercial anode graphite. The time-dependent study suggested a self-sacrificial templating formation mechanism that yielded such intriguing interconnected hollow structures. X-ray photoelectron spectroscopy (XPS) measurements on cycled electrodes indicated that both the deep oxidation of Ni2+ into Ni3+ and the reversible reactions in HCO3- accounted for the ultrahigh capacity of Ni(HCO3)2 in comparison to its generally accepted theoretical capacity of 297 mAh g-1. Morphological characterizations revealed that the interconnected hollow structures enabled the enhanced rate performance and cycling stability due to their larger contact areas with electrolyte and better buffering effect to accommodate the volume change compared to the solid counterpart.

TOC GRAPHICS

ACS Paragon Plus Environment

2

Page 3 of 13

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

ACS Energy Letters

Lithium ion batteries (LIBs) have been widely used in rechargeable energy storage devices as they possess high energy density, long lifespan and benign environmental impact. However, LIBs cannot meet the essential requirements of large-scale high-energy density applications in electric vehicles and smart grids due to the low capacity of commercial electrodes (e.g. graphite, 372 mAh g-1).1-2 The development of new lithium storage materials with high capacity at low cost has emerged as a growing area of research in recent decades.3-4 Transition metal oxides (MxOy, M = Fe, Co, Mn and Ni) have been widely recognized as advanced anodes of LIBs owing to their high theoretical capacities.5-7 Recently, compared to MxOy, transition metal carbonates (MCO3, M = Fe, Co and Mn) have attracted considerable attention as more promising anode materials for lithium storage because of the easier preparation route, lower cost and higher theoretical capacity.8-12 The traditional theoretical capacity (C) of MCO3 (M = Fe, Co or Mn) is calculated to be ~460 mAh g-1 based on the reversible reaction of MCO3 + 2Li+ +2e-  Li2CO3 + M where C = nF/M (n is the number of charge transfer, F is the Faraday constant and M is the molecular weight of the active material). However, most of the reported MCO3 materials exhibit actual capacities over the generally accepted theoretical capacity.13-17 The further reduction of Li2CO3 into lower valence carboncontaining materials (i.e., LixC2, x = 0, 1 or 2) catalyzed by in-situ generated transition metal M nanocrystals has been proved to account for the ultrahigh capacity of MCO3.18-20 The decomposition of Li2CO3 under the catalysis of Ni has been confirmed by the Raman, X-ray diffraction (XRD), and Fourier transform infrared (FTIR) measurements on cycled NiO-Li2CO3 nanocomposites as LIB anode.21 Including the reaction in Li2CO3, that is, Li2CO3 + (4+0.5x)Li+ + (4+0.5x)e-  3Li2O + 0.5LixC2 (x = 0, 1 or 2), the new theoretical capacity of MCO3 (M= Fe, Co or Mn) should be as high as ~1600 mAh g-1 (e.g., n=7 at x=2).11,18-21 Nevertheless, most of the reported MCO3 solid materials display unsatisfactory capacities and cycling stabilities due to the low conductivity and unstable structure. Until now, Ni(HCO3)2 has not yet been reported as an electrode of LIBs despite that NiO is ACS Paragon Plus Environment

3

ACS Energy Letters

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

Page 4 of 13

largely studied as a high-capacity lithium storage material with a theoretical capacity of ~718 mAh g1 22,23

.

Intrigued by the potential ultrahigh theoretical capacity (i.e., 1780 mAh g-1) based on the two

reactions described above that Ni2+ + 2Li  Ni + 2Li+ (n = 2) and 2C4+ + (8 + x)Li  LixC2 + 8Li+ (n = 10 at x = 2) and the excellent supercapacitive properties of Ni(HCO3)2, we developed a facile hydrothermal route to phase-pure interconnected Ni(HCO3)2 hollow spheres.18-21, 24 As an anode of LIBs, they delivered 80th reversible capacities of 1442, 1295 and 1055 mAh g-1 at the current rates (Crates) of 100, 200 and 1000 mA g-1, respectively. The ultrahigh capacity of Ni(HCO3)2 hollow spheres can be attributed to the deep oxidation of Ni2+ into Ni3+ and the reversible reactions in HCO3- as suggested by XPS studies on cycled electrodes, and the outstanding cycling stabiltiy can be attributed to the stable interconnected hollow structure in contrast to the solid counterpart.

Figure 1. (a, b, c) SEM images, (d, e) TEM images and (f) XRD pattern of the synthesized interconnected Ni(HCO3)2 hollow spheres. SEM and TEM images in Figures 1a-e and S1 shows that the 15-h hydrothermal reaction of nickel acetate and ammonium carbonate in the presence of ascorbic acid (AA) at 180oC led to the formation of nanoparticles assembled interconnected hollow spherical aggregates with an average shell thickness of 100 ± 20 nm and diameter of 0.9 ± 0.2 μm. All spheres have a hollow interior and interconnect with

ACS Paragon Plus Environment

4

Page 5 of 13

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

ACS Energy Letters

the adjacent ones as clearly evidenced by the TEM images (Figures 1d and S1b). The powder XRD pattern (Figure 1f) of the hollow spheres presents a pure crystallographic feature of cubic Ni(HCO3)2 crystals (JCPDS No. 15-0782). The average particle size calculated by Scherrer equation based on the full width at half maximum (FWHM) of (110)-face reflection is estimated to be ~27.5 nm for the primary building blocks of Ni(HCO3)2 hollow spheres, which is similar to the actual size (30 ± 3 nm) of nanoparticles from hollow spheres shown in Figure 1e.

Figure 2. (a-d) SEM and TEM (inset) images and (e) XRD patterns of intermediate precipitates collected at the hydrothermal reaction intervals of 1, 2, 5 and 10 h. (f) Schematic illustration of the formation of interconnected Ni(HCO3)2 hollow spheres enabled by employing amorphous Nicontaining spheres as self-sacrificing templates. To reveal the mechanism of forming Ni(HCO3)2 hollow spheres, the time-dependent experiments were conducted (Figures 2, S2 and S3). After reaction for 1 h, randomly aggregated amorphous Nicontaining solid spheres with a smooth surface were obtained (Figures 2a, S2a, S3a and 2e). Then, the solution-mediated dissolution-recrystallization and simultaneous heterogeneous nucleation of Ni(HCO3)2 nanoparticles took place on the surface of amorphous spheres from 2 to 5 h (Figures 2b, 2c, S2b, S2c, S3b and S3c). Finally, due to the gradual dissolution of amorphous pheres from the inside out and the growth of Ni(HCO3)2 shells, crystalline Ni(HCO3)2 hollow spheres were yielded at ACS Paragon Plus Environment

5

ACS Energy Letters

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

Page 6 of 13

10 h (Figures 2d, S2d and S3d). This mechanism also accounts for the gradual increase in crystallinity of Ni(HCO3)2 precipitates as the reaction time increased from 2 to 10 h (Figure 2e). The self-sacrificial templating mechanism is schematically illustrated in Figure 2f. In the absence of AA, the major Ni(HCO3)2 solid spheres with an average diameter of 2.0 ±0.4 µm in conjunction with minor NiO2 and some unknown phases were obtained (Figure S4). This indicates AA exertes a significant influence on the formation of phase-pure Ni(HCO3)2 hollow spheres: the reductive ability of AA prevents the possible oxidation of Ni2+, and the weak acidity of AA allows for the adjustment of pH to promote the dissolution-recrystallization process to produce hollow structures.8,15

Figure 3. Contrastive (a) cycling stabilities and (b) rate performances of Ni(HCO3)2 hollow and solid spheres. Representative (c) voltage profiles operated at 200 mA g-1 and (d) cyclic voltammetry (CV) curves performed at 0.2 mV s-1 of Ni(HCO3)2 hollow spheres. XPS spectra of (e) C1s and (f) Ni2p of different state electrodes of Ni(HCO3)2 hollow spheres. The electrochemical properties of Ni(HCO3)2 hollow and solid spheres were examined using Ni(HCO3)2/Li cell as model system. The initial discharge capacity of hollow spheres was 1668 mAh g-1 with a high initial Coulombic efficiency (CE) of 80.9% operated at 200 mA g-1 (Figure 3a). The capacity fading in the first cycle can be attributed to the formation of solid electrolyte interface (SEI) film and the electrolyte degradation.25 Thereafter, the average CE reached 99.3% (or 99.4%), and an ACS Paragon Plus Environment

6

Page 7 of 13

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

ACS Energy Letters

impressive reversible capacity of 1295 (or 1055) mAh g-1 was maintained at the 80th cycle operated at 200 (or 1000) mA g-1. In contrast, solid spheres showed a rapid capacity decay upon cycling with a low 80th discharge capacity of 162 (or 37) mAh g-1 at 200 (or 1000) mA g-1. As shown in the rate performances (Figure 3b), the hollow spheres exhibited higher CE in each C-rate than that of solid spheres. Even at a high C-rate of 2000 mA g-1, the 50th discharge capacity of hollow spheres still remained 852 mAh g-1, which is ~5.6 times that of solid spheres (151 mAh g-1). When C-rate returned from 2000 to 100 mA g-1, the capacity of hollow (or solid) spheres gradually recovered and reached 1442 (or 638) mAh g-1 at the 80th discharge. Based on the Nitrogen adsorption-desorption and electrochemical impedance spectroscopy measurements (Figure S5), the hollow spheres are advantageous with a higher specific surface area (60 m2 g−1) which can render a closer contact of active anode materials with electrolyte, resulting in higher conductivity (i.e. lower estimated charge-transfer resistance Rct, 75 Ω) than the solid spheres.26-28 As shown in Figure 3c, the voltage profiles from 2nd to 30th cycles almost overlap with each other especaily the charge profiles with three stable plateaus at 1.2, 2.1 and 2.5 V. However, the capacity of the Ni(HCO3)2 solid spheres exhibits rapid decay along with cycles (Figure S6). Figure 3d shows four representative CV profiles at 0.2 mV s-1 with one cathodic peak at 0.88 V but three anodic peaks at 2.55, 2.17 and 1.45 V in the 1st cycle. In contrast, two additional cathodic peaks at 1.66 and 1.35 V appear in the 2nd cycle. Upon the continuous scanning, the position of cathodic peak (1.66 V) gradually decreased into 1.27 V, and the cathodic peak at 0.8 V exhibited stable position but increasing intensity, correlating well with the shifts of voltage plateaus in the discharge profiles and peaks in the corresponding differential capacity vs. voltage (dQ/dV) curves (Figures 3c, S7, S8 and S9). The cells were disassembled at various electrochemical states to examine the cycled Ni(HCO3)2 by XPS (e.g. C1s and Ni2p). As shown in Figures S10a, S10b and 3e, all samples displayed three C1s peaks at the binding energies of ~290, 287 and 285 eV, which can be assigned to HCO3-, C-O and C-C

ACS Paragon Plus Environment

7

ACS Energy Letters

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

Page 8 of 13

bindings, respectively.18 The relative area percentage of each XPS peak in the sample was used to estimate the carbon content of each binding. Figure S10c summarizes the reversible evolution of HCO3- groups in Ni(HCO3)2 during a full cycle. After 10 cycles, the reversible reactions between HCO3- and C-C binding-based species (i.e. LixC2, x=0, 1 or 2) were clearly observed (Figure 3e). The additional reaction in HCO3- could be also proved by the higher capacity of Ni(HCO3)2 than NiO hollow spheres at 1000 mA g-1 (Figure S11). For the Ni2p XPS spectra (Figure 3f), pristine Ni(HCO3)2 hollow spheres showed a single Ni 2p3/2 peak at 856.4 eV, while the 1st charged sample had three Ni 2p3/2 peaks at 850.4, 856.4 and 858.6 eV, which can be attributed to Ni, Ni2+ and Ni3+, respectively.29 For the 10th charged sample, there were only two Ni 2p3/2 peaks at 850.4 eV (Ni) and 858.8 eV (Ni3+). Overall, the contributions from the further oxidation of Ni2+ into Ni3+ and the reversible reactions between HCO3- and LixC2 (x = 0, 1 or 2) should be invoked to calculate the new theoretical capacity of Ni(HCO3)2, yielding 1928 mAh g-1 (n = 13 at x = 2).

Figure 4. (a, b) SEM images and optical micrographs (lower left insets) of the 10th discharged electrodes of Ni(HCO3)2 (a) hollow and (b) solid spheres operated at 200 mA g-1. (c, d) Statistical analysis of diameter variations and (e) corresponding average diameter change percentages compared to the pristine samples of Ni(HCO3)2 (c) hollow and (d) solid spheres in the 10th cycle. (f) Schematic illustration of the volume-buffering and conductivity-enhancement of interconnected Ni(HCO3)2 hollow spheres (upper panels) and the discharge-charge processes resulting electrode disintegration of Ni(HCO3)2 solid spheres (low panels). ACS Paragon Plus Environment

8

Page 9 of 13

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

ACS Energy Letters

Even though Ni(HCO3)2 has an ultrahigh potential theoretical capacity as discussed above, the solid spheres remained much lower capacities after 80 cycles compared to the hollow spheres (Figure 3a). Thus, the critical roles of interconnected hollow structures in the superior lithium storage capability of Ni(HCO3)2 were systematically studied. The SEM and optical images of the 10th cycled electrodes of hollow and solid spheres operated at 200 mA g-1 are shown in Figures 4a, 4b and S12. After the 10th discharge, the hollow spheres are still intact except the increase in both the diameter from pristine 0.9 ±0.2 μm to 1.8±0.3 μm and the wall thickness from pristine 100 ± 20 nm to 360 ± 30 nm (Figures 4a, 4c, S1c-d and S13), but the average diameter of solid spheres are expanded seriously from 2.0 ± 0.4 μm to 4.4 ± 2.3 μm (Figures 4b, 4d and S4f). Conversely, after the 10th charge, the hollow spheres are still intact except the diameter decreases to 1.5± 0.2 μm and the wall thickness reduces to 310 ± 30 nm (Figures 4c, S12a and S13). However, the solid spheres are clearly broken with an average diameter of 2.7 ± 0.7 μm (Figures S12b and 4d). As shown in the percentage changes in the diameter of cycled electrodes (Figure 4e), the solid spheres show about 1.9 times volume shrink (80%) in comparison with that of hollow spheres (43%). Cross-sectional view SEM images of the pristine and 10th cycled electrodes shows that the electrode containing solid spheres experiences approximately 6.1 times thickness shrink (85%) compared to that of hollow spheres (14%) in the 10th cycle (Figures S14 and S15). As shown in the inset of optical images (Figures 4a and S12a), due to the better buffering effect to accommodate the volume change, the cycled electrodes of hollow spheres do not show obvious cracks on their surfaces. However, the cycled electrodes of solid spheres exhibite severe cracks on their surfaces due to the large volume changes in both the particle sizes and electrode thicknesses (Figures 4b and S12b). The cracks of electrodes composed of solid spheres would cause the loss of electrical contact between active material and current collector (copper foil), leading to drastic capacity loss of the cells.30 The advantages of interconnected hollow structures are illustrated in Figure 4f.

ACS Paragon Plus Environment

9

ACS Energy Letters

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

Page 10 of 13

In summary, phase-pure interconnected Ni(HCO3)2 hollow spheres produced via a self-sacrificial templating approach were exploited for the first time as an anode material for LIBs. Intriguingly, these interconnected hollow spheres exhibited superior electrochemical properties with a high reversible capacity. The interconnected and hollow attributes of these spheres played important roles in the excellent cycling stability due to their enhanced structural stability and lithium ion transfer capability. The XPS measurements of cycled electrodes demonstrated the deep redox reactions in both HCO3- and Ni2+, and a high theoretical capacity of 1928 mAh g-1 can be introduced to account for the high 80th capacity of 1442 mAh g-1. As such, these interconnected Ni(HCO3)2 hollow spheres with the elaborate structural characteristics and high-capacity properties may stand out as promising high-capacity anode materials for LIBs. ASSOCIATED CONTENT Supporting Information. Experimental details; SEM, TEM, XRD and electrochemical performances of Ni(HCO3)2 and NiO; SEM and XPS of different state Ni(HCO3)2 electrodes. AUTHOR INFORMATION * E-mail: [email protected] (Q. Shen). * E-mail: [email protected] (Z. Lin). Notes The authors declare no competing financial interest. ACKNOWLEDGMENT We gratefully acknowledge the financial supports from the National Natural Science Foudation of China (21673131), the Taishan Scholar Project of Shandong Province (ts201511004), the Natural Science Foundation of Shandong Province (ZR2016BM03) and Shandong University (2014JC016),

ACS Paragon Plus Environment

10

Page 11 of 13

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

ACS Energy Letters

and Guangzhou Science Technology and Innovation Commission (No. 2016201604030013). S. Zhao acknowledges the financial support from the China Scholarship Council. REFERENCE (1) Dunn, B.; Kamath, H.; Tarascon, J. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011, 334, 928−935. (2) Armand, M.; Tarascon, J. M. Building Better Batteries. Nature 2008, 451, 652−657. (3) Scrosati, B.; Hassoun, J.; Sun, Y.-K. Lithium Ion Batteries. A Look into the Future. Energy Environ. Sci. 2011, 4, 3287−3295. (4) Wu, S.; Xu, R.; Lu, M.; Ge, R.; Iocozzia, J.; Han, C.; Jiang, B.; Lin, Z. Lithium Ion Batteries: Graphene Containing Nanomaterials for Lithium Ion Batteries. Adv. Energy Mater. 2015, 5, 1500400. (5) Liang, J.; Hu, H.; Park, H.; Ding, S. J.; Paik, U.; Lou, X. Construction of Hybrid Bowl-Like Structures by Anchoring NiO Nanosheets on Flat Carbon Hollow Particles with Enhanced Lithium Storage Properties. Energy Environ. Sci. 2015, 8, 1707−1711. (6) Wang, J.; Zhang, Q.; Li, X.; Zhang, B.; Mai, L.; Zhang, K. Smart Construction of ThreeDimensional Hierarchical Tubular Transition Metal Oxide Core/Shell Heterostructures with HighCapacity and Long-Cycle-Life Lithium Storage. Nano Energy 2015, 12, 437−446. (7) Jiang, B.; Han, C.; Li, B.; He, Y.; Lin, Z. In-Situ Crafting of ZnFe2O4 Nanoparticles Impregnated within Continuous Carbon Network as Advanced Anode Materials. ACS Nano 2016, 10, 2728−2735. (8) Zhao, S.; Wei, S.; Liu, R.; Wang, Y.; Yu, Y.; Shen, Q. Cobalt Carbonate Dumbbells for HighCapacity Lithium storage: A Slight Doping of Ascorbic Acid and an Enhancement in Electrochemical Performances. J. Power Sources 2015, 284, 154−161. (9) Zhao, J.; Wang, Y. High-Capacity Full Lithium-Ion Cells Based on Nanoarchitectured Ternary Manganese-Nickel-Cobalt Carbonate and its Lithiated Derivative. J. Mater. Chem. A 2014, 2, 14947−14956. (10) Aragon, M.; Leon, B.; Vicente, C.; Tirado, J. A New Form of Manganese Carbonate for the Negative Electrode of Lithium-Ion Batteries. J. Power Sources 2011, 196, 2863−2866. (11) Garakani, M.; Abouali, S.; Zhang, B.; Takagi, C.; Xu, Z.; Huang, J.; Kim, J. Cobalt Carbonate/and Cobalt Oxide/Graphene Aerogel Composite Anodes for High Performance Li-Ion Batteries. ACS Appl. Mater. Interfaces 2014, 6, 18971−18980. (12) Kang, W.; Yu, D. Y. W.; Li, W.; Zhang, Z.; Yang, X.; Ng, T.-W.; Zou, R.; Tang, Y.; Zhang, W.; Lee, C.-S. Nanostructured Porous Manganese Carbonate Spheres with Capacitive Effects on the High Lithium Storage Capability. Nanoscale 2015, 7, 10146−10151. (13) Zhao, S.; Feng, F.; Yu, F.; Shen, Q. Flower-to-Petal Structural Conversion and Enhanced Interfacial Storage Capability of Hydrothermally Crystallized MnCO3 via the In Situ Mixing of Graphene Oxide. J. Mater. Chem. A 2015, 3, 24095−24102. (14) Gao, M.; Cui, X.; Wang, R.; Wang, T.; Chen, W. Graphene-Wrapped Mesoporous MnCO3 Single Crystals Synthesized by a Dynamic Floating Electrodeposition Method for High Performance LithiumIon Storage. J. Mater. Chem. A 2015, 3, 14126−14133.

ACS Paragon Plus Environment

11

ACS Energy Letters

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

Page 12 of 13

(15) Zhao, S.; Yu, Y.; Wei, S.; Wang, Y.; Zhao, C.; Liu, R.; Shen, Q. Hydrothermal Synthesis and Potential Applicability of Rhombohedral Siderite as a High-Capacity Anode Material for Lithium Ion Batteries. J. Power Sources 2014, 253, 251−255. (16) Zhao, S.; Wang, Y.; Liu, R.; Yu, Y.; Wei, S.; Yu, F.; Shen, Q. Full-Molar-Ratio Synthesis and Enhanced Lithium Storage Properties of CoxFe1-xCO3 Composites with an Integrated Lattice Structure and an Atomic-Scale Synergistic Effect. J. Mater. Chem. A 2015, 3, 17181−17189. (17) Zhong, Y.; Yang, M.; Zhou, X.; Luo, Y.; Wei, J.; Zhou, Z. Orderly Packed Anodes for HighPower Lithium-Ion Batteries with Super-Long Cycle Life: Rational Design of MnCO3/Large-Area Graphene Composites. Adv. Mater. 2015, 27, 806−812. (18) Su, L.; Zhou, Z.; Qin, X.; Tang, Q.; Wu, D.; Shen, P. CoCO3 Submicrocube/Graphene Composites with High Lithium Storage Capability. Nano Energy 2013, 2, 276−282. (19) Ding, Z.; Yao, B.; Feng, J.; Zhang, J. Enhanced Rate Performance and Cycling Stability of a CoCO3-Polypyrrole Composite for Lithium Ion Battery Anodes. J. Mater. Chem. A 2013, 1, 11200−11209. (20) Tian, N.; Hua, C.; Wang, Z.; Chen, L. Reversible Reduction of Li2CO3. J. Mater. Chem. A 2015, 3, 14173−14177. (21) Wang, R.; Yu, X.; Bai, J.; Li, H.; Huang, X.; Chen, L.; Yang, X. Electrochemical Decomposition of Li2CO3 in NiO-Li2CO3 Nanocomposite Thin Film and Powder Electrodes. J. Power Sources 2012, 218, 113−118. (22) Fan, Z.; Liang, J.; Yu, W.; Ding, S.; Cheng, S.; Yang, G.; Wang, Y.; Xi, Y.; Xi, K.; Kumar, R. Ultrathin NiO Nanosheets Anchored on a Highly Ordered Nanostructured Carbon as an Enhanced Anode Material for Lithium Ion Batteries. Nano Energy 2015, 16, 152−162. (23) Wu, J.; Yin, W. J.; Liu, W. W.; Guo, P.; Liu, G.; Liu, X.; Geng, D.; Lau, W. M.; Liu, H.; Liu, L. M. High Performance NiO Nanosheets Anchored on Three-Dimensional Nitrogen-Doped Carbon Nanotubes as a Binder-Free Anode for Lithium Ion Batteries. J. Mater. Chem. A 2016, 4, 10940−10947. (24) Zang, X.; Dai, Z.; Guo, J.; Dong, Q.; Yang, J.; Huang, W.; Dong, X. Controllable Synthesis of Triangular Ni (HCO3)2 Nanosheets for Supercapacitor. Nano Res. 2016, 9, 1358−1365. (25) Ein-Eli, Y. A New Perspective on the Formation and Structure of the Solid Electrolyte Interface at the Graphite Anode of Li-Ion Cells. Electrochem. Solid-State Lett., 1999, 2, 212−214. (26) Wang, J.; Tang, H.; Zhang, L.; Ren, H.; Yu, R.; Jin, Q.; Qi, J.; Mao, D.; Yang, M.; Wang, Y.; et al. Multi-Shelled Metal Oxides Prepared via an Anion-Adsorption Mechanism for Lithium-Ion Batteries. Nature Energy 2016, 1, 16050. (27) Zou, F.; Chen, Y. M.; Liu, K.; Yu, Z.; Liang, W.; Bhaway, S. M.; Gao, M.; Zhu, Y. Metal Organic Frameworks Derived Hierarchical Hollow NiO/Ni/Graphene Composites for Lithium and Sodium Storage. ACS Nano 2016, 10, 377−386. (28) Zhang, L.; Wu, H.; Madhavi, S.; Hng, H.; Lou, X. Formation of Fe2O3 Microboxes with Hierarchical Shell Structures from Metal-Organic Frameworks and Their Lithium Storage Properties. J. Am. Chem. Soc. 2012, 134, 17388−17391. (29) Watanabe, N.; Arakawa, T.; Sasaki, Y.; Yamashita, T.; Koiwa, I. Influence of the Memory Effect on X-Ray Photoelectron Spectroscopy and Raman Scattering in Positive Electrode of Ni-MH Batteries. J. Electrochem. Soc. 2012, 159, A1949−A1953.

ACS Paragon Plus Environment

12

Page 13 of 13

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

ACS Energy Letters

(30) Ko, M.; Chae, S.; Ma, J.; Kim, N.; Lee, H. -W.; Cui, Y.; Cho, J. Scalable Synthesis of SiliconNanolayer-Embedded Graphite for High-Energy Lithium-Ion Batteries. Nature Energy 2016, 1, 16113.

ACS Paragon Plus Environment

13