Hollow Mesoporous Co(PO3)2@Carbon Polyhedra as High

5 days ago - School of Materials Science & Engineering, Nanyang Technological ... National Institute of Education, Nanyang Technological University , ...
0 downloads 0 Views 5MB Size
Article pubs.acs.org/JPCC

Cite This: J. Phys. Chem. C XXXX, XXX, XXX−XXX

Hollow Mesoporous Co(PO3)2@Carbon Polyhedra as High Performance Anode Materials for Lithium Ion Batteries Shiji Hao,†,‡ Bo Ouyang,§ Chaojiang Li,† Bowei Zhang,† Jianyong Feng,† Junsheng Wu,*,∥ Madhavi Srinivasan,*,† and Yizhong Huang*,†,‡ †

College of Science, Hubei University of Technology, Hongshan District, Wuhan 430068, China School of Materials Science & Engineering, Nanyang Technological University, Singapore 639798, Singapore § National Institute of Education, Nanyang Technological University, 1 Nanyang Walk, Singapore 637616, Singapore ∥ Institute for Advanced Materials and Technology, University of Science and Technology Beijing, 30 Xueyuan Road, Beijing 100083, China

J. Phys. Chem. C Downloaded from pubs.acs.org by UNIV OF LOUISIANA AT LAFAYETTE on 03/27/19. For personal use only.



S Supporting Information *

ABSTRACT: The hollow mesoporous Co(PO3)2@carbon nanocomposite (H−Co(PO3)2@C) was synthesized using ZIF-67 as the template by a facile one-step thermal decomposition reaction. As an anode for lithium ion batteries, its reversible capacity remains up to 601 mAh g−1 at 1 C after 500 cycles. Such a high reversible capacity along with the excellent rate capability and long-term cycling stability benefits from the hollow mesoporous structure and uniform carbon framework encapsulated active nanocrystals. These results render the as-prepared H−Co(PO3)2@C to be a promising anode material for high performance lithium ion batteries.

1. INTRODUCTION Lithium ion batteries (LIBs) have been utilized in various portable electronic devices and shown great potential as electrochemical storage devices for electric vehicles (EVs), hybrid electric vehicles (HEVs), and plug-in HEVs.1−3 However, the limitation of commercial LIBs’ capacity is emerging due to the dramatic development and increased power consumption of electronic products. Therefore, there exists an urgent demand to find and develop novel materials to replace traditional carbon-based anodes for LIBs which have a very low capacity of 372 mAh g−1.4−6 Recently, various Cobased compounds, such as Co3O4, CoS2, and CoP, have been intensively studied and investigated as anode materials for LIBs due to their high theoretical capacities (1000−600 mAh g−1) and appealing promise.7−11 According to the proposed reaction mechanism from Poizot et al., their reversible electrochemical behavior involves the formation and decomposition of lithium compounds accompanied by the reduction and oxidation of metal species in LIBs.12−14 The overall reaction progress of Co3O4 with lithium is

Based on the previous studies, the presence of substitution elements, such as P and B atoms acting as buffer matrix, could effective eliminate nanoparticle polymerization, satisfying the long-term stability requirement for lithium intercalation/ deintercalation.19−23 Nevertheless, these materials still show a high irreversible capacity and poor cyclic stability, ascribing to their low conductivity and unaccepted mechanical properties.15,20−22,24,25 Numerous studies have demonstrated that the remarkable improvement of the anode materials’ structural/electrochemical properties can be achieved by constructing a variety of hybrid nanostructures.26−32 For example, Wu et al. synthesized hollow structured cobalt sulfide@porous carbon polyhedra/carbon nanotubes composites, which shows a high performance as anode materials for LIBs (1668 mAh g−1 after 100 cycles).33 Wang et al. prepared an onionlike carbon matrix supported Co3O4 nanocomposite with an extremely high rate capability at various current densities that exhibits an improved battery performance (632 mAh g−1 at 200 mA g−1).34 The designable hybrid nanostructures, which have well-defined conductive frameworks and reserved inner space, could not only facilitate the access of electrolyte and reduce diffusion

Co3O4 + 8Li → 3Co + 4Li 2O

However, the cycling stability of the electrodes is still limited due to the degradation of active materials. During repetitive cycling, the large volume change (around 100%) and aggregation of metallic atoms are inevitable, leading to the disintegration of electrodes and rapid loss of capacity.15−18 © XXXX American Chemical Society

Received: December 28, 2018 Revised: March 4, 2019

A

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

battery test system (Neware, BTS-5 V10 mA, China). The cyclic voltammetry (CV) test and electrochemical impedance (EIS) measurement (10 mHz−100 kHz, amplitude of 10 mV) were conducted via the electrochemical workstation (autolab, PGSTAT302N).

pathway but also accommodate the huge volume change induced stains and stresses.35−40 Herein, we propose a facile method to prepare a new hybrid nanostructure composed of Co(PO3)2 nanocrystals embedded in hollow mesoporous carbon polyhedra through the thermal decomposition and the simultaneous combination reaction of ZIF-67 templates. The introduced conductive carbon matrix and hollow porous nanostructure are expected to increase the electronic conductivity, facilitate the diffusion of Li ions, and alleviate the volume change during electrochemical cycling. Benefiting from the constructed hybrid nanostructure, the asprepared nanocomposite is exploited as an anode material for LIBs and expected to achieve superior electrochemical properties.

3. RESULTS AND DISCUSSION As a well-studied template, ZIF-67 shows the structure of zeolitic imidazolate framework (ZIF) formed by the coordination of 2-methyl imidazolate ligands and Co cations and possesses exceptional chemical and thermal stability.42−44 Benefitting from its unique topology and thermal behavior, the well-defined Co(PO3)2@C polyhedra were fabricated through tunable annealing conditions, and the control of their internal structure was demonstrated. Figure 1 illustrates

2. EXPERIMENTAL SECTION 2.1. Materials Synthesis. All the chemicals were purchased from Sigma-Aldrich and used in analytical grades without further purification. The template of ZIF-67 was synthesized via a general chemical deposition method, as reported in Sun et al.’s work.41 First of all, methanol solutions of cobalt nitrate and 2-methylimidazole solution with the concentrations of 40 and 160 mmol L−1, respectively, were prepared. Two solutions were then thoroughly mixed. After 24 h, the as-obtained ZIF-67 nanoparticles were collected by centrifugation of the mixture, followed by washing several times with methanol, and finally dried in a vacuum oven. To prepare Co(PO3)2@C nanocomposites, the as-obtained ZIF-67 particles were completely mixed with red phosphorus at a weight ratio of ZIF-67/P of 1:2 through grinding for at least 10 min. The mixture was subsequently annealed at 600 °C for 5 h with a nitrogen flow. The ramping rate during thermal treatment was tuned to control the nanostructure of the prepared nanocomposites. The H−Co(PO3)2@C and solid Co(PO3)2@carbon (S−Co(PO3)2@C) were generated at a heating rate of 1 and 10 °C min−1, respectively. 2.2. Characterization. X-ray diffraction (XRD) patterns were recorded on Bruker D8 with Cu Kα radiation. The nanostructure and morphology of products were revealed by field-emission scanning electron microscopy (FESEM, JEOL7600) and transmission electron microscopy (TEM, JEOL 2100). Micro-Raman measurements of samples were performed using a WiTec Alpha300 system with a 532 nm wavelength incident laser light. The specific surface areas and pore structure were analyzed using the Micromeritics ASAP Tristar II 3020. The thermogravimetric analysis (TGA) was conducted on the TA Instruments DMA Q900. X-ray photoelectron spectroscopy (XPS, Omicrometer analyzer EA 125) was used to analyze the surface element electron state. 2.3. Electrochemical Measurement. The electrochemical evaluation as anode materials for LIBs was performed using CR2016-type coin cells. To prepare working electrode, the slurry of 70% active materials, 20% Super P carbon, and 10% poly(vinylidene fluoride) (PVDF) was coated on copper foil, before the electrode was remained for drying overnight at 60 °C. The coin cells were assembled in a glovebox (filled with argon, moisture, and oxygen level ≤1 ppm). The 1 M LiPF6 of ethylene carbonate (EC) and diethyl carbonate (DEC) (1:1 by volume) solution was utilized as an electrolyte. The pure lithium foils and Celgard 2400 membrane were used as counter electrodes and the separator, respectively. The charge− discharge measurement was determined at room temperature (25 °C) in the voltage range of 0.01−3.0 V, using the Neware

Figure 1. Synthetic scheme for the preparation of S− and H− Co(PO3)2@C.

schematically the preparation of Co(PO3)2@C nanocomposites. First, rhombic dodecahedral ZIF-67 particles with a highly uniform size of 500 nm were synthesized (Figure S1a,b). The transformation from ZIF-67 to Co(PO3)2@C polyhedra was subsequently performed via the phosphating process at high temperature. The measured and simulated X-ray diffraction (XRD) patterns of the formed ZIF-67 are shown in Figure S1c. All the XRD peaks in the range of 5°−50° match well with the simulated result, indicating the phase purity of the as-prepared template. The morphologies of the as-prepared H−Co(PO3)2@C and S−Co(PO3)2@C were revealed by SEM measurement. Compared with the original ZIF-67 template, H−Co(PO3)2@C exhibits a slight surface shrinkage due to the pyrolysis of organic species, but the pristine precursor shape with an average size of 500 nm is still maintained (Figure 2a,b). In contrast, S−Co(PO3)2@C appears more severe surface contraction (Figure 2c,d). The rough surface can be attributed to the high thermal annealing ramping rate, leading to the slightly reduced particle size. Figure 3 shows two XRD patterns of the as-prepared H−Co(PO3)2@C and S−Co(PO3)2@C particles. The diffraction peaks can be indexed as monoclinic Co(PO3)2 (PDF card no. 00-027-1120, I*/a, a = 1.12 nm, b = 0.829 nm, c = 0.993 nm), suggesting the successful phosphating process during the thermal annealing treatment. Compared with the XRD pattern of H−Co(PO3)2@C, S−Co(PO3)2@C generates much broader XRD peaks, indicating its nanocrystalline structure. This ultrafine size is apparently the result of the higher heating rate. X-ray photoelectron spectroscopy (XPS) was further conducted to analyze the chemical state of the as-prepared samples present in the near surface region. In the highB

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 2. Low and high magnified FESEM images of H−Co(PO3)2@C (a, b) and S−Co(PO3)2@C (c, d).

2p3/2 and Co 2p1/2, at binding energies of approximately 782.9 and 799.0 eV, indicating the 2+ oxidation state of Co transition metal in the present metaphosphate system.45−47 The peak at around 134.4 eV in the P 2p energy level of H−Co(PO3)2@C (upper panel in Figure 4b) can be ascribed to the P in the metaphosphate group and confirm the existence of PO3−.40,48,49 In addition, the O 1s spectrum (upper panel in Figure 4c) is deconvoluted into two peaks 531.9 and 533.2 eV, which were assigned to the nonbridging oxygen atom (P−O− Co) and bridging oxygen atoms (P−O−P), respectively.48,50 In comparison with the XPS spectra (Figure 4a−c) obtained on H− and S−Co(PO3)2@C, it was revealed that the binding energies in Co, P, and O regions were slightly downshifted by 0.3, 0.2, and 0.2 eV, respectively, after the heating rate was increased to 10 °C min−1. It would contribute to the reduced electronic interaction between Co, P, and O, affecting materials’ activity as anodes for lithium ion batteries.40,51 The detailed structure and crystallinity of the as-prepared H−Co(PO3)2@C were characterized via transmission electron microscopy (TEM) analysis (Figure 5). A low-magnification TEM image is shown in Figure 5a, where the clear contrast between the interior and the edges of the particle shell suggests

Figure 3. XRD patterns of the as-prepared H−Co(PO3)2@C and S− Co(PO3)2@C with standard PDF card of Co(PO3)2.

resolution XPS spectrum of Figure 4a (upper panel), Co 2p regions of H−Co(PO3)2@C show spin−orbit components, Co

Figure 4. High-resolution XPS spectra of Co 2p (a), P 2p (b), and O 1s (c) collected on H−Co(PO3)2@C (upper) and S−Co(PO3)2@C (lower). C

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 5. TEM images (a, b), HRTEM image (c), and EDX elemental mapping (d) of the as-prepared H−Co(PO3)2@C.

using N2 adsorption data. Both as-prepared composites show a broad mesoporous distribution centered at around 20 nm, which is probably attributed to the void spaces between polyhedra particles. The Raman measurement was conducted to investigate the carbon structure of the obtained composites. As shown in Figure 6, two peaks are visible at about 1377 and 1571 cm−1,

the internal cavity of the polyhedra composite. As seen in the magnified TEM image (Figure 5b), the hollow structure consists of numerous Co(PO3)2 nanoparticles (black dots) with size of around 20 nm that are homogeneously embedded in porous carbon matrix (gray matrix). Figure 5c is a highresolution TEM (HRTEM) image of the Co(PO3)2 nanoparticles, which are seen to be encapsulated into the hollow carbon framework. According to the fast Fourier transformation (FFT) pattern along the [−120] zone axis (inset), the distinct lattice fringes are calculated to have a d-spacing of 0.42 nm corresponding to the (002) lattice plane of Co(PO3)2. The elemental mapping was conducted to reveal the elemental distribution. As indicated in Figure 5d, Co, P, O, and C elements within the as-prepared polyhedra composite are present and homogeneously distributed in consistence with the TEM result. As compared (Figure S2a−d), when the heating ramp was tuned to be 10 °C min−1, the obvious concaves from facets and edges are observed, in good agreement with SEM results. In addition, the as-prepared S−Co(PO 3) 2@C, composed of pure phase Co(PO3)2 crystals and carbon framework, shows solid nanostructure and appears cookielike. The S−Co(PO3)2@C composite exhibits smaller diameter of Co(PO3)2 particles (∼5 nm), which is in alignment with the XRD measurement. The nitrogen (N2) adsorption−desorption isotherms and pore size distribution curves of the as-obtained H−Co(PO3)2@C and S−Co(PO3)2@C composites, derived from Brunauer−Emmett−Teller (BET) measurement, are presented in Figure S3a,b. The N2 adsorption−desorption of H− Co(PO3)2@C composite can be identified as type IV with a typical H3 hysteresis loop, according to the IUPAC classification (Figure S3a). It suggests the coexistence of meso- and macropores. After the heating rate was set to be 10 °C min−1, an isotherms of type II is observed for the S− Co(PO3)2@C particles (Figure S3b), which is a typical characteristic of nonporous materials, in good agreement with the TEM image (Figure S2a−d). The respective BET specific surface areas of H−Co(PO3)2@C and S−Co(PO3)2@ C composites are calculated to be 13.4 and 7.4 m2 g−1. The higher specific surface area of H−Co(PO3)2@C is contributed by the interior hollow cavity and mesoporous framework and offers the enhanced electrolyte contact area and ionic diffusion rate, leading to an excellent electrochemical performance. The pore size distribution curves (the inset in Figure S3a,b) were obtained from the Barrett−Joyner−Halenda (BJH) method

Figure 6. Raman spectra of the as-prepared H−Co(PO3)2@C and S− Co(PO3)2@C.

which can be assigned to D and G bands, respectively, indicating the existence of carbon materials. For the H− Co(PO3)2@C composite obtained via annealing at 1 °C min−1, the calculated peak ratio of the IG/ID intensity is 0.91, which is 2 times higher than 0.42 for the S−Co(PO3)2@C, suggesting an improved degree of graphitic content in carbon framework. It is a result of long-term high-temperature carbonization, which further improves the electronic conductivity of the encapsulated Co(PO3)2 nanoparticles. The other peaks located at around 916 and 1100 cm−1 are typical characteristic of Co(PO3)2.52 As the heating rate was raised up to 10 °C min−1, Raman bands for Co(PO3)2 become less prominent and are slightly shifted due to the confinement effect, indicating the size reduction of crystals.53 To quantify the amount of Co(PO3)2 in the as-synthesized H−Co(PO3)2@C product, thermogravimetric analysis (TGA) was performed in air at a rate of 10 °C min−1. As shown in Figure S4, the initial weight loss (16.0%) from room temperature to 250 °C is assigned to the removal of absorbed and coordinated water. The following major weight loss takes place in the temperature ranging from D

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

Figure 7. CV curves at a scan rate of 0.5 mV s−1 (a), discharge and charge profiles (b), the cycling performance at 0.2 C (c), and the rate capability (d) of the H−Co(PO3)2@C electrode; the cycling performance (e) of the H−Co(PO3)2@C electrode at 1 C; Nyquist plots (f) of the H− Co(PO3)2@C and S−Co(PO3)2@C electrodes after 50 cycles.

300 to 700 °C, which is due to the decomposition of carbon species. The mass ratio of Co(PO3)2 is calculated to be 69.2% based on the final weight retention. To evaluate the electrochemical performance of the assynthesized composites as anodes for LIBs, a series of electrochemical measurements were performed. Figure 7a presents CV curves of the first five cycles of the H− Co(PO3)2@C electrode at a scan rate of 0.5 mV s−1 in the potential window between 0.01 and 3.00 V. On the basis of previous studies, the strong cathodic peak at 0.81 V in the first cycle can be assigned to the irreversible decomposition of Co(PO3)2 and the formation of SEI film and Li2O matrix (Figure 7a).10,54 This peak disappears in the following cycle, confirming the occurrence of textural modification during lithiation/delithiation process. Meanwhile, the two pairs of cathodic and anodic peaks at 0.56, 1.34 V and 0.01, 0.45 V suggest a two-step electrochemical redox reaction related to the reversible formation of CoxP. Compared to the S− Co(PO3)2@C electrode (Figure S5a), the H−Co(PO3)2@C electrode shows a better overlapping of the subsequent cycling curves, indicating its better stability. In addition, the lithiation

process of carbon species is identified from the peak around 0.01 V. The detailed electrochemical reactions above can be written as the following equations: Co(PO3)2 + 6Li+ + 6e− → CoP2 + 3Li 2O CoP2 + 6Li+ + 6e− ↔ Co + 2Li3P

C + x Li + x e− ↔ LixC

The electrochemical performance of the H−Co(PO3)2@C composite was further examined by galvanostatic discharge− charge tests, which were conducted at a current density of 0.2 C (1 C = 1000 mA g−1) in the potential range from 0.01 to 3.0 V. As seen in Figure 7b, the discharge−charge profile shows poorly defined discharge plateau regions and a steep charge voltage plateau in the initial cycle. The potential shifts are observed in the subsequent discharge process, which is consistent with the CV result. The initial discharge and charge capacities are 1085.6 and 695.7 mAh g−1, respectively, corresponding to an initial Coulombic efficiency of 64.1%. The low initial efficiency may be attributed to the irreversible capacity loss, including inevitable formation of SEI and Li2O E

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C

electronic and ionic conductivities, leading to significant improvement on the cycling and rate performance. After 500 cycles at 1 C, the coin cell of the H−Co(PO3)2@ C was disassembled in a glovebox. The as-obtained active materials cast from the electrode were further characterized by TEM to reveal its morphology evolution during cycling. As seen in Figure S6a, the H−Co(PO3)2@C still maintains the initial morphologies even after 500 cycles under a high current density, claiming its significant electrochemical stability. Besides, the agglomeration of Co(PO3)2 particles during electrochemical processes is successfully eliminated, and the crystal size is shown to be around 30 nm in Figure S6b. Herein, we can conclude that the excellent electrochemical performance of the H−Co(PO3)2@C electrode is mainly contributed by the hollow and porous structured carbon matrix. The porous conductive matrix with inner void effectively increases the available active sites resulting in both higher electronic conductivity and larger lithium storage. Meanwhile, the increased contact surface area provides the short path length of lithium ion diffusion and facilitates fast cycling rates.

and the incomplete extraction of lithium from the active material. The reversible capacity (695.7 mAh g−1) is much higher than the calculated theoretical value of 600 mAh g−1 for Co(PO3)2. The extra capacity can be contributed from the two parts. The polymeric/gel-like films, on the one hand, are reversibly formed/dissolved on the surface of active materials, corresponding to the decomposition of electrolyte.55 On the other hand, through the initial lithiation reaction, the fresh metallic Co surface area is generated, leading to an electrocatalysis process resulting in the further reduction of carbon species in SEI film along with formation of Li2O and the chemical energy transfer for energy storage.56 After 50 cycles (Figure 7c), the discharge capacity slowly decreases and stabilizes at 517.0 mAh g−1, showing the retention rate of 74.3%. For comparison, the S−Co(PO3)2@C composite was also investigated under the same conditions. As illustrated in Figure S5b,c, the S−Co(PO3)2@C composite shows a poor cycling performance. After 50 cycles, the reversible capacity gradually fades down to a low value of 461.2 mAh g−1, which effectively proves the significant merits from the hollow interior and mesoporous nanostructure of the H−Co(PO3)2@C composite. The rate capability of the H−Co(PO3)2@C at progressively increased current densities is shown in Figure 7d. It is clearly observed that the as-formed H−Co(PO3)2@C electrode possesses great capacity retention under different current densities. Even through the current density is increased to 10 C, the specific capacity is still maintained up to 305 mAh g−1, which is much higher than the capacity of the S−Co(PO3)2@ C composite (Figure S5d). The reversible capacities are achieved to be 658.2, 578.5, 524.2, 469.3 and 385.3 at current rates of 0.2, 0.5, 1, 2, and 5 C, respectively. Moreover, when the current rate is dropped back to 0.2 C, the specific capacity of the H−Co(PO3)2@C electrode recovers to 612.5 mAh g−1. To further confirm the superior high rate properties of the hollow structured composite, the cycling performance of the as-fabricated H−Co(PO3)2@C electrode was evaluated at a high rate of 1 C for 500 cycles. As shown in Figure 7e, the reversible capacity of the H−Co(PO3)2@C is attained at 601.0 mAh g−1 after 500 cycles. It is noteworthy that the discharge capacity rapidly drops down to 511 mAh g−1 within 32 cycles and gradually rises back to around 600 mAh g−1 during the subsequent cycling. Similar phenomena have been widely observed in transition metal compound based anode materials, originated from the kinetically reversible growth of the polymeric gel-like film induced by the activated electrolyte decomposition and the refinement of the hollow and porous nanostructure initiating the optimization of stable SEI film.55,57 To further clarify the excellent electrochemical performance of the H−Co(PO3)2@C particles, the EIS measurements were carried out. Figure 7f shows the drawn Nyquist plots of the H− and S−Co(PO3)2@C anodes after 50 cycles. The EIS spectra consist of a semicircle and a straight line corresponding to the limitations in the charge transfer reaction (Rct) and the lithium ion diffusion, respectively. In the charge separation domain, the much smaller diameter for the H−Co(PO3)2@C implies its low charge transfer resistance compared to that of the S− Co(PO3)2@C. In addition, the H−Co(PO3)2@C electrode shows a more vertical straight line at the low-frequency region, confirming its faster ionic diffusion upon cycling. The result demonstrates that the constructed hollow interior and porous nanostructure allow the effective increase of active materials’

4. CONCLUSIONS H−Co(PO3)2@C composites as anode materials for LIBs were fabricated via a thermal decomposition process using ZIF-67 as the template. The resultant composite possesses a unique nanostructure consisting of Co(PO3)2 particles encapsulated in hollow and mesoporous structured carbon frameworks. The asobtained morphology not only relaxes the stresses and strains of Li+ insertion/extraction and increases the active surface area but also prevents the aggregation of Co(PO3)2 and enhances the electronic and ionic conductivity. As a result, the H− Co(PO3)2@C anode exhibits a high specific capacity of 601 mAh g−1 under the current density of 1 C over 500 cycles. Considering its superior performance along with the simple and cost-effective synthesis method, the H−Co(PO3)2@C will be one of the promising anode candidates for LIBs with high energy density.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.8b12494. SEM images and XRD pattern of ZIF-67 template, TEM images and EDX elemental mapping of S−Co(PO3)2@ C, nitrogen sorption isotherms and pore size distribution curves of H− and S−Co(PO3)2@C, TGA curve of H− Co(PO 3)2 @C, TEM images of H−Co(PO 3)2 @C particles after cycles, electrochemical data of S−Co(PO3)2@C anode (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (Y.H.). *E-mail: [email protected] (M.S.). *E-mail: [email protected] (J.W.). ORCID

Yizhong Huang: 0000-0003-2644-856X Notes

The authors declare no competing financial interest. F

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C



Rate and Long-Life Lithium Ion Batteries. J. Phys. Chem. C 2014, 118, 2263−2272. (18) Han, F.; Zhang, C. Z.; Sun, B.; Tang, W.; Yang, J. X.; Li, X. K. Dual-Carbon Phase-Protective Cobalt Sulfide Nanoparticles with Cable-Type and Mesoporous Nanostructure for Enhanced Cycling Stability in Sodium and Lithium Ion Batteries. Carbon 2017, 118, 731−742. (19) Kim, E.; Son, D.; Kim, T. G.; Cho, J.; Park, B.; Ryu, K. S.; Chang, S. H. A Mesoporous/Crystalline Composite Material Containing Tin Phosphate for Use as the Anode in Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2004, 43, 5987−5990. (20) Li, Y.; Li, J. Carbon-Coated Macroporous Sn2P2O7 as Anode Materials for Li-Ion Battery. J. Phys. Chem. C 2008, 112, 14216− 14219. (21) Lee, J. Y.; Xiao, Y.; Liu, Z. Amorphous Sn2P2O7, Sn2B2O5 and Sn2BPO6 Anodes for Lithium Ion Batteries. Solid State Ionics 2000, 133, 25−35. (22) Li, L.; Zhang, J.; Zou, Y.; Jiang, W.; Lei, W.; Ma, Z. High-Rate and Long-Term Cycle Stability of Lithium-Ion Batteries Enabled by Boron-Doping TiO2 Nanofiber Anodes. J. Electroanal. Chem. 2019, 833, 573. (23) Sun, C.-F.; Hu, J.; Wang, P.; Cheng, X.-Y.; Lee, S. B.; Wang, Y. Li3PO4 Matrix Enables a Long Cycle Life and High Energy Efficiency Bismuth-Based Battery. Nano Lett. 2016, 16, 5875−5882. (24) Xiao, Y.; Lee, J.; Yu, A.; Liu, Z. Electrochemical Performance of Amorphous and Crystalline Sn2P2O7 Anodes in Secondary Lithium Batteries. J. Electrochem. Soc. 1999, 146, 3623−3629. (25) Hu, L.; Zheng, S.; Chen, Z.; Huang, B.; Yang, J.; Chen, Q. 3D Graphene Modified Sphere-Like VPO4/C as a High-Performance Anode Material for Lithium-Ion Batteries. Electrochim. Acta 2018, 284, 609−617. (26) Cao, S.; Feng, X.; Song, Y.; Liu, H.; Miao, M.; Fang, J.; Shi, L. In Situ Carbonized Cellulose-Based Hybrid Film as Flexible Paper Anode for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 1073−1079. (27) Shi, L.; Pang, C.; Chen, S.; Wang, M.; Wang, K.; Tan, Z.; Gao, P.; Ren, J.; Huang, Y.; Peng, H.; Liu, Z. Vertical Graphene Growth on SiO Microparticles for Stable Lithium Ion Battery Anodes. Nano Lett. 2017, 17, 3681. (28) Wu, S.; Xia, T.; Wang, J.; Lu, F.; Xu, C.; Zhang, X.; Huo, L.; Zhao, H. Ultrathin Mesoporous Co3O4 Nanosheets-Constructed Hierarchical Clusters as High Rate Capability and Long Life Anode Materials for Lithium-Ion Batteries. Appl. Surf. Sci. 2017, 406, 46−55. (29) Lee, K. S.; Park, S.; Lee, W.; Yoon, Y. S. Hollow Nanobarrels of α-Fe2O3 on Reduced Graphene Oxide as High-Performance Anode for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 2027− 2034. (30) Wu, C.; Kopold, P.; van Aken, P. A.; Maier, J.; Yu, Y. High Performance Graphene/Ni2P Hybrid Anodes for Lithium and Sodium Storage through 3D Yolk−Shell-Like Nanostructural Design. Adv. Mater. 2017, 29, 1604015. (31) Chen, C.; Xie, X.; Anasori, B.; Sarycheva, A.; Makaryan, T.; Zhao, M.; Urbankowski, P.; Miao, L.; Jiang, J.; Gogotsi, Y. MoS2-OnMxene Heterostructures as Highly Reversible Anode Materials for Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2018, 57, 1846−1850. (32) Wang, Y.; Liu, P.; Zhu, K.; Wang, J.; Yan, K.; Liu, J. One-Step Fabrication of in Situ Carbon-Coated NiCo2O4@ C Bilayered Hybrid Nanostructural Arrays as Free-Standing Anode for High-Performance Lithium-Ion Batteries. Electrochim. Acta 2018, 273, 1−9. (33) Wu, R.; Wang, D. P.; Rui, X.; Liu, B.; Zhou, K.; Law, A. W.; Yan, Q.; Wei, J.; Chen, Z. In-Situ Formation of Hollow Hybrids Composed of Cobalt Sulfides Embedded within Porous Carbon Polyhedra/Carbon Nanotubes for High-Performance Lithium-Ion Batteries. Adv. Mater. 2015, 27, 3038−3044. (34) Wang, Y.; Yan, F.; Liu, S. W.; Tan, A. Y. S.; Song, H.; Sun, X. W.; Yang, H. Y. Onion-Like Carbon Matrix Supported Co3O4 Nanocomposites: A Highly Reversible Anode Material for Lithium Ion Batteries with Excellent Cycling Stability. J. Mater. Chem. A 2013, 1, 5212−5216.

ACKNOWLEDGMENTS This research was supported by Tier 1 (AcRF Grant MOE Singapore M4011528, M4011648, and M4011959), Natural Science Fund of China (51762023), and the Chinese Natural Science Foundation (Grants 60906053, 62174118, and 51308050309).



REFERENCES

(1) Li, W.; Yang, Z.; Li, M.; Jiang, Y.; Wei, X.; Zhong, X.; Gu, L.; Yu, Y. Amorphous Red Phosphorus Embedded in Highly Ordered Mesoporous Carbon with Superior Lithium and Sodium Storage Capacity. Nano Lett. 2016, 16, 1546−1553. (2) Lu, J.; Chen, Z.; Ma, Z.; Pan, F.; Curtiss, L. A.; Amine, K. The Role of Nanotechnology in the Development of Battery Materials for Electric Vehicles. Nat. Nanotechnol. 2016, 11, 1031−1038. (3) Ma, J.; Hu, P.; Cui, G.; Chen, L. Surface and Interface Issues in Spinel LiNi0.5Mn1.5O4: Insights into a Potential Cathode Material for High Energy Density Lithium Ion Batteries. Chem. Mater. 2016, 28, 3578−3606. (4) Li, Y.; Zhu, C.; Lu, T.; Guo, Z.; Zhang, D.; Ma, J.; Zhu, S. Simple Fabrication of a Fe2O3/Carbon Composite for Use in a HighPerformance Lithium Ion Battery. Carbon 2013, 52, 565−573. (5) Dong, Y.; Ma, R.; Hu, M.; Cheng, H.; Lee, J.-M.; Li, Y. Y.; Zapien, J. A. Polymer-Pyrolysis Assisted Synthesis of Vanadium Trioxide and Carbon Nanocomposites as High Performance Anode Materials for Lithium-Ion Batteries. J. Power Sources 2014, 261, 184− 187. (6) Chen, L.; Zhou, G.; Liu, Z.; Ma, X.; Chen, J.; Zhang, Z.; Ma, X.; Li, F.; Cheng, H. M.; Ren, W. Scalable Clean Exfoliation of HighQuality Few-Layer Black Phosphorus for a Flexible Lithium Ion Battery. Adv. Mater. 2016, 28, 510−517. (7) Hao, Q.; Zhao, D.; Duan, H.; Xu, C. Porous Co3O4/CuO Composite Assembled from Nanosheets as High-Performance Anodes for Lithium-Ion Batteries. ChemSusChem 2015, 8, 1435− 1441. (8) Sun, J.; Liu, H.; Chen, X.; Evans, D. G.; Yang, W. An Oil Droplet Template Method for the Synthesis of Hierarchical Structured Co3O4/C Anodes for Li-Ion Batteries. Nanoscale 2013, 5, 7564− 7571. (9) Chen, W.; Li, T.; Hu, Q.; Li, C.; Guo, H. Hierarchical CoS2@C Hollow Microspheres Constructed by Nanosheets with Superior Lithium Storage. J. Power Sources 2015, 286, 159−165. (10) Yang, D.; Zhu, J.; Rui, X.; Tan, H.; Cai, R.; Hoster, H. E.; Yu, D. Y.; Hng, H. H.; Yan, Q. Synthesis of Cobalt Phosphides and Their Application as Anodes for Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2013, 5, 1093−1099. (11) Lu, A.; Zhang, X.; Chen, Y.; Xie, Q.; Qi, Q.; Ma, Y.; Peng, D.-L. Synthesis of Co2P/Graphene Nanocomposites and Their Enhanced Properties as Anode Materials for Lithium Ion Batteries. J. Power Sources 2015, 295, 329−335. (12) Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J. Nano-Sized Transition-Metal Oxides as Negative-Electrode Materials for Lithium-Ion Batteries. Nature 2000, 407, 496. (13) Li, W.-Y.; Xu, L.-N.; Chen, J. Co3o4 Nanomaterials in LithiumIon Batteries and Gas Sensors. Adv. Funct. Mater. 2005, 15, 851−857. (14) Yuan, Z.; Huang, F.; Feng, C.; Sun, J.; Zhou, Y. Synthesis and Electrochemical Performance of Nanosized Co3O4. Mater. Chem. Phys. 2003, 79, 1−4. (15) Yang, Y.; Wang, B.; Zhu, J.; Zhou, J.; Xu, Z.; Fan, L.; Zhu, J.; Podila, R.; Rao, A. M.; Lu, B. Bacteria Absorption-Based Mn2P2O7− Carbon@ Reduced Graphene Oxides for High-Performance LithiumIon Battery Anodes. ACS Nano 2016, 10, 5516−5524. (16) Li, L.; Zhou, G.; Shan, X.-Y.; Pei, S.; Li, F.; Cheng, H.-M. Co3O4 Mesoporous Nanostructures@ Graphene Membrane as an Integrated Anode for Long-Life Lithium-Ion Batteries. J. Power Sources 2014, 255, 52−58. (17) Sun, H.; Sun, X.; Hu, T.; Yu, M.; Lu, F.; Lian, J. GrapheneWrapped Mesoporous Cobalt Oxide Hollow Spheres Anode for HighG

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C (35) Liu, N.; Lu, Z.; Zhao, J.; McDowell, M. T.; Lee, H.-W.; Zhao, W.; Cui, Y. A Pomegranate-Inspired Nanoscale Design for LargeVolume-Change Lithium Battery Anodes. Nat. Nanotechnol. 2014, 9, 187. (36) Wu, H.; Yu, G.; Pan, L.; Liu, N.; McDowell, M. T.; Bao, Z.; Cui, Y. Stable Li-Ion Battery Anodes by in-Situ Polymerization of Conducting Hydrogel to Conformally Coat Silicon Nanoparticles. Nat. Commun. 2013, 4, 1943. (37) Zheng, F.; Yang, Y.; Chen, Q. High Lithium Anodic Performance of Highly Nitrogen-Doped Porous Carbon Prepared from a Metal-Organic Framework. Nat. Commun. 2014, 5, 5261. (38) Hu, L.; Liu, N.; Eskilsson, M.; Zheng, G.; McDonough, J.; Wågberg, L.; Cui, Y. Silicon-Conductive Nanopaper for Li-Ion Batteries. Nano Energy 2013, 2, 138−145. (39) Wang, G.; Wen, Z.; Yang, Y.-E.; Yin, J.; Kong, W.; Li, S.; Sun, J.; Ji, S. Ultra-Long Life Si@ RGO/g-C3N4 with a Multiply Synergetic Effect as an Anode Material for Lithium-Ion Batteries. J. Mater. Chem. A 2018, 6, 7557−7565. (40) Huang, J.; Sun, Y.; Zhang, Y.; Zou, G.; Yan, C.; Cong, S.; Lei, T.; Dai, X.; Guo, J.; Lu, R. J. A. M.; Li; Xiong. A New Member of Electrocatalysts Based on Nickel Metaphosphate Nanocrystals for Efficient Water Oxidation. Adv. Mater. 2018, 30, 1705045. (41) Sun, C.; Yang, J.; Rui, X.; Zhang, W.; Yan, Q.; Chen, P.; Huo, F.; Huang, W.; Dong, X. MOF-Directed Templating Synthesis of a Porous Multicomponent Dodecahedron with Hollow Interiors for Enhanced Lithium-Ion Battery Anodes. J. Mater. Chem. A 2015, 3, 8483−8488. (42) Jiang, Z.; Li, Z.; Qin, Z.; Sun, H.; Jiao, X.; Chen, D. Ldh Nanocages Synthesized with MOF Templates and Their High Performance as Supercapacitors. Nanoscale 2013, 5, 11770−11775. (43) Wu, R.; Qian, X.; Rui, X.; Liu, H.; Yadian, B.; Zhou, K.; Wei, J.; Yan, Q.; Feng, X. Q.; Long, Y.; Wang, L.; Huang, Y. Zeolitic Imidazolate Framework 67-Derived High Symmetric Porous Co3O4 Hollow Dodecahedra with Highly Enhanced Lithium Storage Capability. Small 2014, 10, 1932−1938. (44) Yu, D.; Wu, B.; Ge, L.; Wu, L.; Wang, H.; Xu, T. Decorating Nanoporous Zif-67-Derived NiCo2O4 Shells on a Co3O4 Nanowire Array Core for Battery-Type Electrodes with Enhanced Energy Storage Performance. J. Mater. Chem. A 2016, 4, 10878−10884. (45) Gond, R.; Rao, R. P.; Pralong, V.; Lebedev, O. I.; Adams, S.; Barpanda, P. J. I. c. Cubic Sodium Cobalt Metaphosphate [NaCo(PO3)3] as a Cathode Material for Sodium Ion Batteries. Inorg. Chem. 2018, 57, 6324−6332. (46) Tan, L.; Luo, Z.; Liu, H.; Yu, Y. J. J. o. A. Compounds, Synthesis of Novel High-Voltage Cathode Material LiCoPO4 Via Rheological Phase Method. J. Alloys Compd. 2010, 502, 407−410. (47) Zhan, Y.; Lu, M.; Yang, S.; Xu, C.; Liu, Z.; Lee, J. Y. J. C. Activity of Transition-Metal (Manganese, Iron, Cobalt, and Nickel) Phosphates for Oxygen Electrocatalysis in Alkaline Solution. ChemCatChem 2016, 8, 372−379. (48) Liu, Q.; Chen, C.; Zheng, J.; Wang, L.; Yang, Z.; Yang, W. 3D Hierarchical Ni (PO3)2 Nanosheet Arrays with Superior Electrochemical Capacitance Behavior. J. Mater. Chem. A 2017, 5, 1421− 1427. (49) Ahn, H. S.; Tilley, T. D. J. A. F. M. Electrocatalytic Water Oxidation at Neutral PH by a Nanostructured Co(PO3)2 Anode. Adv. Funct. Mater. 2013, 23, 227−233. (50) Tsuchida, J.; Schneider, J.; Rinke, M. T.; Eckert, H. Structure of Ternary Aluminum Metaphosphate Glasses. J. Phys. Chem. C 2011, 115, 21927−21941. (51) Xu, C.; Sun, J.; Gao, L. J. J. o. M. C. Direct Growth of Monodisperse SnO2 Nanorods on Graphene as High Capacity Anode Materials for Lithium Ion Batteries. J. Mater. Chem. 2012, 22, 975− 979. (52) Ahn, H. S.; Tilley, T. D. Electrocatalytic Water Oxidation at Neutral PH by a Nanostructured Co(PO3)2 Anode. Adv. Funct. Mater. 2013, 23, 227−233. (53) Zi, J.; Büscher, H.; Falter, C.; Ludwig, W.; Zhang, K.; Xie, X. Raman Shifts in Si Nanocrystals. Appl. Phys. Lett. 1996, 69, 200−202.

(54) Choi, C.; Seo, S.-D.; Shim, H.-W.; Dar, M. A.; Cho, I. S.; Kim, D.-W. Facile Synthesis and Electroactivity of 3-D Hierarchically Superstructured Cobalt Orthophosphate for Lithium-Ion Batteries. J. Alloys Compd. 2015, 652, 100−105. (55) Laruelle, S.; Grugeon, S.; Poizot, P.; Dolle, M.; Dupont, L.; Tarascon, J. On the Origin of the Extra Electrochemical Capacity Displayed by Mo/Li Cells at Low Potential. J. Electrochem. Soc. 2002, 149, A627−A634. (56) 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. (57) Sun, H.; Xin, G.; Hu, T.; Yu, M.; Shao, D.; Sun, X.; Lian, J. High-Rate Lithiation-Induced Reactivation of Mesoporous Hollow Spheres for Long-Lived Lithium-Ion Batteries. Nat. Commun. 2014, 5, 4526.

H

DOI: 10.1021/acs.jpcc.8b12494 J. Phys. Chem. C XXXX, XXX, XXX−XXX