In situ grown Nb4N5 nanocrystal on nitrogen-doped graphene as a novel anode for lithium ion battery

Chenlong Donga, Xin Wanga, Xiangye Liua, Xiaotao Yuana, Wujie Donga, Houlei Cuib, Yuhang Duanc and Fuqiang Huang*ab
aState Key Laboratory of Rare Earth Materials Chemistry and Applications, Beijing National Laboratory for Molecular Sciences, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871, P. R. China. E-mail: huangfq@mail.sic.ac.cn; huangfq@pku.edu.cn
bCAS Key Laboratory of Materials for Energy Conversion, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, P. R. China
cFaculty of Science, Applied Chemistry Program, Beijing University of Chemical Technology, Beijing 100029, P. R. China

Received 26th May 2016 , Accepted 20th August 2016

First published on 22nd August 2016


Abstract

The metal-rich niobium nitride of Nb4N5 has higher conductivity than Nb3N5 and a higher theoretical specific capacity than NbN. To rationally design a metal-rich anode material, Nb4N5 nanocrystals coated by nitrogen-doped graphene (N-G) have been successfully synthesized by a facile in situ ice bathing method with subsequent annealing in NH3. The use of these as an anode material is reported for the first time. The discharge capacity is 487 mA h g−1 at the current density of 0.1 A g−1 (0.0819 mA cm−2) after 200 cycles and the high rate discharge capacity is 125 mA h g−1 at a current density of 5 A g−1 (4.0926 mA cm−2). Specially, the discharge capacity is still enhanced after 200 cycles at 0.1 A g−1 (0.0819 mA cm−2). The Nb4N5/N-G hybrid could be a promising anode material for LIBs with a high rate performance and long cycle life.


Introduction

Lithium ion batteries (LIBs) have attracted considerable attention in the field of energy storage and conversion. The performance of LIBs is mainly determined by the properties of active electrode materials which are controlled by composition, structure, morphology and so on.1 Commercial LIBs usually use graphite as an anode. However, the theoretical capacity of graphite is severely limited to 372 mA h g−1, which limits its application for electrical/hybrid vehicles.2–4 Generally, many metal oxides have been explored as high capacity anode materials, however, they suffer from low conductivity. Compared to metal oxides, the metal nitrides often show improved conductivity. Therefore, the nitridation of metal oxides is a suitable method to acquire high performance anode materials.

Metal nitrides are emerging as promising electrode materials for high-performance LIBs5,6 due to their high Li+ diffusion, excellent electrical conductivity7 and flat potential close to that of lithium metal.8 Niobium nitrides are mainly composed of Nb3N5, Nb4N3, NbN and some other nonstoichiometric niobium nitride compounds. Among them, the metal-rich phases (such as Nb4N5 and Nb5N6) have also been presented in δ-NbN system, which comprise lots of Nb5+ ions.9,10 They are promising candidates for LIBs because of their high-span valency transformation.10

Herein, we report the synthesis of Nb4N5/nitrogen-doped graphene hybrid nanomaterial (Nb4N5/N-G) by in situ growth in ice bathing followed by NH3 annealing. The as-prepared Nb4N5/N-doped graphene hybrid is reported for the first time using it as a LIBs anode material, which exhibits excellent electrochemical performance. Besides superior conductivity of Nb4N5, the internal defects between graphene layers motivated by the substitution of nitrogen atoms may enhance lithium storage.11 Meanwhile, nitrogen-doped graphene can remit the expansion of Nb4N5 availably during discharge process and help forming stable solid electrolyte interface (SEI) layer. Therefore, the Nb4N5/nitrogen-doped graphene hybrid can be a novel promising anode for LIBs.

Experimental section

Material

Niobium pentachloride was purchased from Aladdin Chemical Reagent. Anhydrous ethanol (>99.7%) was purchased from Beijing Chemical Works. Ammonium hydroxide was purchased from Beijing Tong Guang Fine Chemicals Company. Graphene oxide was prepared according to the previous reported literature by Hummers12 from graphite powder (purchased from Aldrich, powder, <20 micron). All used water in the process was deionized.

Material synthesis

Typically, 0.5489 g NbCl5 was dissolved into 10 mL ethanol to form solution A. 90 g ice was added into 30 mL 3 mg mL−1 GO solution (solution B) by vigorously magnetic stirring. The solution A was added into solution B dropwise. Then, the pH of mixed solution was adjusted to ca. 7 by ammonium hydroxide. After centrifuging, the resulting Nb2O5/GO precipitate was washed by water for three times and freeze-dried for 12 h. Finally, the Nb2O5/GO was converted to Nb4N5/N-G by annealing in NH3 at 700 °C under the flow rate of 350 mL min−1 for 4 h.

For comparison, the fabrication of Nb4N5 was similar to the above.

Physical characterization

The microstructures were studied by scanning electron microscope (SEM) using a ZEISS Merlin Compact Field-emission scanning electron microscope (ZEISS, Germany) and transmission electron microscopy (TEM) used a JEM-2100 electron microscope (JEOL Ltd, Japan) working at 200 kV. The surface chemical composition was studied by X-ray Photoelectron Spectroscopy using Axis Ultra Imaging Photoelectron Spectrometer (Kratos Analytical Ltd, Japan). X-ray diffraction (XRD) was performed on a D8 Focus diffractometer with monochromatized Cu Kα radiation (λ = 1.5418 Å). Raman spectrum was measured using a Renishawi instrument via Raman Microscope with laser excitation at 532 nm. Nitrogen adsorption isotherms were measured using a Micromeritics ASAP 2020 analyzer at 196 °C. The weight of the sample used in the measurement was 100–200 mg. The Brunauer–Emmett–Teller (BET) method was utilized to calculate the specific surface area.

Electrochemical tests

To prepare the working electrodes, the Nb4N5/N-G, Nb4N5 and N-G were mixed with carbon black (Super P) and polyvinylidene fluoride binder with a ratio of 8[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 in N-methyl-2-pyrrolidone (NMP) to form slurry, respectively. The three slurries were coated onto copper foil. The coated foil was cut into disk electrodes of 14 mm in diameter. Then the electrodes were vacuum-dried overnight at 120 °C. Lithium foil (China Energy Lithium Co., Ltd) was used as the counter and reference electrode. A glass fiber from Whatman was used as separator. The electrolyte was 1 M LiPF6 in a 50[thin space (1/6-em)]:[thin space (1/6-em)]50 (w/w) mixture of ethylene carbonate (EC) and dimethyl carbonate (DMC). Coin Cell (LIR-2016 type) assembly was carried out in a recirculating an Argon glovebox where both the moisture and oxygen contents were below 1 ppm. All the cells were cycled between 0.01 and 3 V (vs. Li/Li+) at certain current densities on a LAND CT2001A battery test system. The rate capability was evaluated by varying the current density from 0.1 A g−1 to 5 A g−1. Specific capacities were calculated based on the mass of the each activated materials. Cyclic voltammetry (CV) was obtained on a CHI1000C electrochemistry workstation at the scan rate of 0.3 mV s−1 and the potential from 0.01 to 3 V (vs. Li+/Li). The electrochemical impedance spectrum was obtained on the CHI600E with amplitude of 10 mV in the frequency range from 100 kHz to 0.01 Hz.

Results and discussion

Characterization

XRD patterns of Nb4N5, N-G and Nb4N5/N-G are shown in Fig. 1. The as-synthesized pure N-G displays a broad diffraction peak of (002) at 27.5° which is assigned to wrinkle layer-to-layer distance.13 The diffraction peaks of pure Nb4N5 at 36.1°, 41.7°, 60.9°, 73.2°, 76.4° are indexed to (211), (310), (312), (213) and (422) crystal lattices, respectively (JCPDS No. 51-1327, I4/m). The Nb4N5/N-G diffraction peak curve is higher than pure Nb4N5 from 25° to 35° because of the peak overlay of N-G and Nb4N5. Then, the Nb4N5/N-G hybrid is confirmed by Raman spectroscopy. As shown in Fig. 2b, two broad peaks emerge at 1340 and 1590 cm−1 which are attributed to D and G bands of nitrogen-doped graphene, respectively. The D band strongly associates with the disorder degree of graphene.14,15 The G band corresponds to the zone center E2g mode related to phonon vibrations in sp2 carbon materials.16,17 Clearly, the ID/IG ratio (1.1[thin space (1/6-em)]:[thin space (1/6-em)]1) indicates the presence of a large amount of defects in the graphene layer due to the doping of nitrogen atoms. The Raman spectrum reveals that nitrogen-doped graphene with crystalline structures and many defects are obtained, which are favorable for the electrochemical properties.18
image file: c6ra13647h-f1.tif
Fig. 1 (a) XRD patterns of Nb4N5, N-G and Nb4N5/N-G hybrid; (b) Raman spectrum of Nb4N5/N-G hybrid.

image file: c6ra13647h-f2.tif
Fig. 2 (a) SEM image of Nb4N5/N-G hybrid; (b) low-magnification TEM image of Nb4N5/N-G hybrid; (c) high-magnification TEM image of Nb4N5/N-G hybrid. (Note: △ and ★ stand for N-G and Nb4N5, respectively.)

The morphology of Nb4N5/N-G hybrid is characterized by SEM and TEM, as shown in Fig. 2. The typical wrinkle structure of N-G is clearly observed in Fig. 2a, which results from thermodynamically spontaneous bending. Meanwhile, the nanocrystals of Nb4N5 are tightly captured by the N-G. Fig. 2b further confirms that the nanocrystals of Nb4N5 are uniformly dispersed on N-G without severe aggregation because of the protection of N-G. The SAED pattern (Fig. 3c) can further confirm the high crystallinity of Nb4N5/N-G. These diffraction rings can be indexed to (312) of Nb4N5, (002) of N-G and (220) of Nb4N5 which are well in accordance with XRD patterns (Fig. 1a).


image file: c6ra13647h-f3.tif
Fig. 3 (a) XPS survey. High-resolution Nb4N5/N-G XPS spectrum of (b) Nb 3d, (c) C 1s and (d) N 1s. The black scatter is the raw date and the red curve is the sum of different fitting peaks.

X-ray photoelectron spectroscopy (XPS) measurement is performed to study the surface chemical composition and oxidation state. As shown in Fig. 3a, the XPS survey spectrum shows that there are four elements: C, N, Nb and O without other impurity. Nb 3d spectrum involves two distinct doublets and a weak shoulder peak (Fig. 3a). The highest Nb 3d5/2 binding energy at 207.0 eV can be indexed to the Nb5+–O due to the oxidation of surface. The lowest binding energy emerges at 203.8 eV which belongs to the Nb 3d5/2 in Nb4N5 and the peak located at 205.0 eV can be indexed to Nb5+–N.19–21 The peaks at 206.3, 208.0 and 209.9 eV correspond to Nb 3d3/2. The C 1s spectrum is displayed in Fig. 3b, which mainly exhibits three conspicuous peaks. The obvious core level peak binding energy at 284.9 eV represents graphite-like sp2 C–sp2 C, indicating that the most parts of carbon atoms exist in conjugated lattice.18 The relatively weak peaks binding energy at 285.7 and 286.6 eV can be ascribed to N–sp2 C and N–sp3 C, respectively, because of N atoms and the defects.22 The results of N–sp2 C and N–sp3 C correspond with the Raman analysis. As shown in Fig. 3d, the broad N 1s XPS peak can be fitted to the five peaks of Nb–N (396.9 eV), pyridinic N (398.5 eV), pyrrolic N (399.8 eV), graphitic N (401.5 eV) and oxygenated N (402.8 eV).

To investigate the content of N-G in Nb4N5/N-G, the TG-DSC measurement was carried out. As shown in Fig. 4, the weight increase from ca. 150 to ca. 400 °C because the 1 molecular Nb4N5 is oxidated to be 2 molecular Nb2O5. After then, the weight decreases from ca. 400 to ca. 800 °C due to the combustion of N-G and the oxidation of Nb4N5. The content of N-G is calculated in detail (see in ESI). According to the analysis and calculation, the content of N-G in Nb4N5/N-G is about 16.78%.


image file: c6ra13647h-f4.tif
Fig. 4 The TG/DSC curves of Nb4N5/N-G.

Three kinds of materials are examined by N2 adsorption–desorption measurements (Fig. S1). The BET surface area of bare Nb4N5 is 44.53 m2 g−1 and the N-G is 36.86 m2 g−1. After the Nb4N5 is anchored on N-G, the dispersity of Nb4N5 loading on N-G is enhanced, so that the BET surface area increases into 49.64 m2 g−1.

Electrochemical performance

In Nb4N5/N-G, Li ion can intercalate first and later on under conversion reaction. During the process, a large number of Li+ can be stored at a low potential for lithium ion battery anode material. The reaction of Nb4N5 and Li can be speculated as follow:23
Nb4N5 + xLi → LixNb4N5 (intercalation reaction)

LixNb4N5 + (15 − x)Li → 4Nb + 5Li3N (conversion reaction)

The cycle voltammogram (Fig. 5a) presents a stable electrochemical performance of Nb4N5/N-G. During the first cathodic process, two apparent cathodic peaks are observed at 1.41 V and 0.50 V (vs. Li/Li+). The origin of cathodic peak at 1.41 V can be attributed to the Li ion intercalation into the Nb4N5/N-G. The peak at 0.50 V is due to the solid electrolyte interface (SEI) formation. Furthermore, the minor slope changing at 0.15 V can be assigned to a conversion reaction from LixNb4N5 to Nb and Li3N. After the 5th cycle, the peak at 1.41 V shifts to a higher potential (1.58 V) due to the difficulty of Li intercalation into residual Nb4N5/N-G. Similarly, the peak at 0.15 V shifts to a prominent peak at 0.17 V because of the electrode polarization. During the anodic sweep, the oxidation peak can be attributed to the oxidation reaction, from Nb0 to LixNb4N5 and LixNb4N5 to Nb4N5.


image file: c6ra13647h-f5.tif
Fig. 5 (a) Cyclic voltammogram of Nb4N5/N-G hybrid with a scanning rate of 0.3 mV s−1 ranging from 0.01 to 3 V (vs. Li+/Li); (b) Nyquist plots of Nb4N5/N-G and Nb4N5 at opening circuit voltage after 3 cycles.

To further clarify the difference of Nb4N5/N-G and Nb4N5 in electrochemical performance, electrochemical impedance spectroscopy (EIS) is carried out at frequencies from 100 kHz to 0.01 Hz to identify the correlation between the electrochemical performance and electrode kinetics. Fig. 5b shows the Nyquist plots for the Nb4N5/N-G and Nb4N5 after three cycles. The Nyquist plots for Nb4N5/N-G and Nb4N5 share the similar feature of a high-frequency depressed semicircle and a medium-frequency depressed semicircle followed by a linear tail in the low-frequency region. The intercept on the Z′ axis at the high frequency end is the equivalent series resistance (Rs). The Nb4N5 and Nb4N5/N-G have a similar intercept on the Z′ axis. The size of the semicircular that includes the medium-frequency response represents charge-transfer resistance (Rct). Obviously, the diameter of the semicircle for Nb4N5/N-G electrode is markedly lower than Nb4N5 in the high-frequency region. This phenomenon indicates that Nb4N5/N-G possesses lowest contact and charge-transfer impedances so that electron can transport rapidly during the electrochemical lithium extraction and insertion.24 These results confirm that the disorderly wrinkle layer-to-layer N-G (likes huge charge-transfer network) can provide more active sites and greater contact area with Nb4N5. The inclined linear line in the low-frequency region stands for the Warburg impedance (Zw) which relates to lithium diffusion in the solid electrode.25 All the advantages of Nb4N5/N-G above result in enormous improvement of the electrochemical performance (specific capacity, rate performance, etc.) compared to bare Nb4N5.

Fig. 6a shows galvanostatic charge profiles of Nb4N5/N-G electrode within the voltage of 0.01–3 V (vs. Li/Li+) at a current density of 0.1 A g−1 (0.0819 mA cm−2). The profiles are in good agreement with the CV results. Two conspicuous plateaus (1.5 and 2.2 V vs. Li/Li+) are obtained during the discharge while prominent plateaus at around 2.1 V is observed in charge process. The galvanostatic charge/discharge curves of Nb4N5/N-G, Nb4N5 and N-G at a current of 0.1 A g−1 over a potential range of 0.01–3.0 V (vs. Li/Li+) are shown in Fig. 6b and c and S2, respectively. At the first cycle, the charge/discharge capacities of Nb4N5, N-G and Nb4N5/N-G are 432.4/899.2, 461.4/771.3 and 663.3/941.8 mA h g−1, respectively. Apparently, the coulombic efficiency gets prodigious increasement from 48.09%, 59.8% to 70.43% which reflects synergistic effect between Nb4N5 and N-G. The enhanced initial coulombic efficiency may be ascribed to the catalytic sites on the surface of N-G, related to the decomposition of electrolyte occupied by Nb4N5.26 This phenomenon effectively helps the formation of solid electrolyte interface (SEI) and the remission of severe volume expansion, leading a decreased irreversible capacity. Fig. 5a shows that the reversible capacity of Nb4N5/N-G retains at 487.3 mA h g−1 after 200 cycles, with no capacity loss from the 20th cycle. Meanwhile, the tendency of charge–discharge curve still maintains enhanced step by step. The metal nitride reveals ordered and disordered phases with significant levels of lithium vacancies.27 Li vacancies are the charge carriers in Li3N, which is a well-known fast-ion conductor with the highest Li+ conductivity.27 The formation of additional vacancies in the nitride phases implies their potential for enhanced lithium ion diffusion. The reason why the Nb4N5/N-G has remarkable performance can be interpreted by following: (1) suitable kinetic reaction due to the fast diffusion of Li+; (2) tremendous conductive network of N-G.


image file: c6ra13647h-f6.tif
Fig. 6 (a) Galvanostatic discharge–charge profiles of Nb4N5/N-G within the voltage window of 3.00–0.01 V at a current density of 0.1 A g−1 (0.0819 mA cm−2); (b) cycle performance and coulombic efficiency of Nb4N5/N-G at 0.1 A g−1 (0.0819 mA cm−2); (c) cyclic performance and coulombic efficiency of Nb4N5 at 0.1 A g−1 (0.0955 mA cm−2).

The rate performance of Nb4N5/N-G with different current densities is given in Fig. 7a. The incremental rates discharge capacity are 481, 430, 365, 265, 125 mA h g−1 at 0.2 A g−1 (0.1637 mA cm−2), 0.5 A g−1 (0.4093 mA cm−2), 1 A g−1 (0.8185 mA cm−2), 2 A g−1 (1.6370 mA cm−2), 5 A g−1 (4.0926 mA cm−2), respectively. After high rate charge/discharge, the specific capacity of Nb4N5/N-G at the current density of 0.1 A g−1 (0.0819 mA cm−2) can recover to the initial value, which confirms the structure stability. The capacity of Nb4N5/N-G hybrid at even 5 A g−1 (4.0926 mA cm−2) is as high as 125 mA h g−1 while that of Nb4N5 (Fig. 7b) is only 70 mA h g−1 at the current density of 1 A g−1 (0.9549 mA cm−2). The N-G provides enormous conductive network for Nb4N5, which is favorable for charge transport. The Nb4N5/N-G hybrid has more active sites because of the metal-rich material and N-G. After discharge, the generated Nb atoms take lower area leading fast Li+ diffusion. The rate performance of Nb4N5/N-G hybrid anode further confirms the possible application for LIBs.


image file: c6ra13647h-f7.tif
Fig. 7 (a) Rate performance of Nb4N5/N-G at varied current density; (b) rate performance of bare Nb4N5 at varied current density.

Conclusions

In conclusion, Nb4N5/N-G hybrid is successfully synthesized by a facile in situ ice bathing method with subsequent annealing in NH3. The conductive Nb4N5 nanocrystals are uniformly dispersed on the N-G. The Nb4N5 nanocrystals have mainly no band gap which make contributions to lower charge-transfer resistance and electrode polarization so that the cells have the excellent electrochemical performance.28 The specific capacity is 487 mA h g−1 at 0.1 A g−1 (0.0819 mA cm−2) which still demonstrates upgrade tendency. The hybrid also exhibits good rate capability compared to the bare Nb4N5 with capacity of 125 mA h g−1 at the current density of 5 A g−1 (4.0926 mA cm−2). Above all, Nb4N5/N-G hybrid can be a novel promising anode material for LIBs with high energy density and long cycle life.

Acknowledgements

This work was financially supported from National Natural Science Foundation of China (Grant no. 61376056, 51125006, 91122034, 51121064, 51222212), National Program of China (Grant No. 2016YFB0901600), and Science and Technology Commission of Shanghai (Grant no. 13JC1405700, 14520722000).

Notes and references

  1. X. Xia, D. Chao, Y. Zhang, Z. Shen and H. Fan, Nano Today, 2014, 9, 785–807 CrossRef CAS.
  2. G. Cui, L. Gu, L. Zhi, N. Kaskhedikar, P. A. van Aken, K. Müllen and J. Maier, Adv. Mater., 2008, 20, 3079–3083 CrossRef CAS.
  3. M. Yoshio, H. Wang, K. Fukuda, T. Umeno, T. Abe and Z. Ogumi, J. Mater. Chem., 2004, 14, 1754–1758 RSC.
  4. H. Li, Z. Wang, L. Chen and X. Huang, Adv. Mater., 2009, 21, 4593 CrossRef CAS.
  5. G. Cui, L. Gu, A. Thomas, L. Fu, P. A. van Aken, M. Antonietti and J. Maier, ChemPhysChem, 2010, 11, 3219–3223 CrossRef CAS PubMed.
  6. N. Suzuki, R. B. Cervera, T. Ohnishi and K. Takada, J. Power Sources, 2013, 231, 186–189 CrossRef CAS.
  7. X. Lu, G. Wang, T. Zhai, M. Yu, S. Xie, Y. Ling, C. Liang, Y. Tong and Y. Li, Nano Lett., 2012, 12, 5376–5381 CrossRef CAS.
  8. B. Das, M. Reddy, P. Malar, T. Osipowicz, G. S. Rao and B. Chowdari, Solid State Ionics, 2009, 180, 1061–1068 CrossRef CAS.
  9. M. Wen, C. Hu, Q. Meng, Z. Zhao, T. An, Y. Su, W. Yu and W. Zheng, J. Phys. D: Appl. Phys., 2009, 42, 035304 CrossRef.
  10. H. Cui, G. Zhu, X. Liu, F. Liu, Y. Xie, C. Yang, T. Lin, H. Gu and F. Huang, Adv. Sci., 2015, 2(12), 1500126 CrossRef.
  11. L. Tang, Y. Wang, Y. Li, H. Feng, J. Lu and J. Li, Adv. Funct. Mater., 2009, 19, 2782–2789 CrossRef CAS.
  12. W. S. Hummers Jr and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef.
  13. D. Geng, S. Yang, Y. Zhang, J. Yang, J. Liu, R. Li, T.-K. Sham, X. Sun, S. Ye and S. Knights, Appl. Surf. Sci., 2011, 257, 9193–9198 CrossRef CAS.
  14. Y. T. Lee, N. S. Kim, S. Y. Bae, J. Park, S.-C. Yu, H. Ryu and H. J. Lee, J. Phys. Chem. B, 2003, 107, 12958–12963 CrossRef CAS.
  15. K. Suenaga, M. Yudasaka, C. Colliex and S. Iijima, Chem. Phys. Lett., 2000, 316, 365–372 CrossRef CAS.
  16. A. Ferrari, J. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. Piscanec, D. Jiang, K. Novoselov and S. Roth, Phys. Rev. Lett., 2006, 97, 187401 CrossRef CAS PubMed.
  17. A. Malesevic, R. Vitchev, K. Schouteden, A. Volodin, L. Zhang, G. Van Tendeloo, A. Vanhulsel and C. Van Haesendonck, Nanotechnology, 2008, 19, 305604 CrossRef PubMed.
  18. H. Wang, C. Zhang, Z. Liu, L. Wang, P. Han, H. Xu, K. Zhang, S. Dong, J. Yao and G. Cui, J. Mater. Chem., 2011, 21, 5430 RSC.
  19. Y. Ufuktepe, A. Farha, S. Kimura, T. Hajiri, K. Imura, M. Mamun, F. Karadag, A. Elmustafa and H. Elsayed-Ali, Thin Solid Films, 2013, 545, 601–607 CrossRef CAS.
  20. Y. Ufuktepe, A. H. Farha, S.-i. Kimura, T. Hajiri, F. Karadağ, M. A. Al Mamun, A. A. Elmustafa, G. Myneni and H. E. Elsayed-Ali, Mater. Chem. Phys., 2013, 141, 393–400 CrossRef CAS.
  21. J. Alfonso, J. Buitrago, J. Torres, J. Marco and B. Santos, J. Mater. Sci., 2010, 45, 5528–5533 CrossRef CAS.
  22. A. L. M. Reddy, A. Srivastava, S. R. Gowda, H. Gullapalli, M. Dubey and P. M. Ajayan, ACS Nano, 2010, 4, 6337–6342 CrossRef CAS PubMed.
  23. D. K. Nandi, U. K. Sen, D. Choudhury, S. Mitra and S. K. Sarkar, ACS Appl. Mater. Interfaces, 2014, 6, 6606–6615 CAS.
  24. C. He, S. Wu, N. Zhao, C. Shi, E. Liu and J. Li, ACS Nano, 2013, 7, 4459–4469 CrossRef CAS PubMed.
  25. J. Z. Wang, C. Zhong, D. Wexler, N. H. Idris, Z. X. Wang, L. Q. Chen and H. K. Liu, Chem.–Eur. J., 2011, 17, 661–667 CrossRef CAS PubMed.
  26. I. s. Kim, P. Kumta and G. Blomgren, Electrochem. Solid-State Lett., 2000, 3, 493–496 CrossRef CAS.
  27. Z. Stoeva, R. Gomez, A. G. Gordon, M. Allan, D. H. Gregory, G. B. Hix and J. J. Titman, J. Am. Chem. Soc., 2004, 126, 4066–4067 CrossRef CAS PubMed.
  28. C. Chen, Y. Huang, H. Zhang, X. Wang, G. Li, Y. Wang, L. Jiao and H. Yuan, J. Power Sources, 2015, 278, 693–702 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Fabrication of graphene oxide (GO), TG-DSC analysis and calculation and coulombic efficiency of N-G at 0.1 A g−1 (0.0832 mA cm−2). See DOI: 10.1039/c6ra13647h

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.