Yajin Haoab,
Zhiqiang Shib,
Jing Wang*b and
Qiang Xu*a
aDepartment of Chemical Engineering, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, P. R. China. E-mail: xuqiang@tju.edu.cn; Fax: +86 22 27401684; Tel: +86 22 27890322
bLaboratory of Fiber Modification and Functional Fiber, School of Materials Science and Engineering, Tianjin Polytechnic University, Tianjin 300387, P. R. China. E-mail: wangjing@tjpu.edu.cn; Fax: +86 22 83955055; Tel: +86 22 83955833
First published on 18th March 2015
Nitrogen doped mesoporous carbon (NMC) with high specific surface area, large pore volume and stable nitrogen content has been prepared by in situ doping through one step carbonization to immobilize sulfur for lithium–sulfur batteries. The structure and composition of the prepared NMC and NMC@S composites are confirmed with nitrogen sorption, elemental analysis, X-ray photoelectron spectroscopy (XPS), X-ray diffraction patterns (XRD), high resolution transmission electron microscopy (HR-TEM) and thermogravimetric analysis (TGA). Scanning electron microscopy (SEM) images show a homogeneous distribution of sulfur within NMC@S. In comparison with free doped mesoporous carbon (MC), the NMC@S with high sulfur content (65%) is found to exhibit a higher initial discharge capacity of 1012 mA h g−1 and higher capacity retention. The electrochemical performance of Li–S batteries is improved by the nitrogen doping, which not only enhances the electric conductivity but also further confines polysufides to the mesoporous carbon from shuttle effect.
Furthermore, the gradual loss of active mass from the cathode into the electrolyte and onto the Li metal anode results in ‘shuttle effection’,3 severe self-discharge, increased-resistance, low coulombic efficiency and fast capacity decay on cycling,4 which hinders the use of Li–S batteries in practical applications.
Hitherto, numerous efforts have been employed to address the polysulfide shuttle challenge in Li–S batteries. Porous carbon with strong conductivity, high specific surface area and large pore volume is introduced into Li–S batteries to immobilize S and its discharge product, ensuring the desired sulfur content as well as high energy density, such as mesoporous carbon,1,5–8 microporous carbon,9,10 carbon nanotubes11,12 and hollow porous carbon.13 The nano-scaled architecture accommodating a nano-sized sulfur composite shows enhanced utilization of the active sulfur, and shortens the distance for ionic and electronic transportation. Simultaneously, sulfur and the polysulfide can be adsorbed to the porous carbon, which refrains the insulated discharge product (Li2S or Li2S2) from block deposition onto the electrode, thus alleviating polarization and extending cycle life of Li–S battery.
Although porous carbon materials with diverse textural characteristics have been designed for sulfur–carbon composite cathodes and shown improved performances, the properties of carbon materials as a matrix are also affected by the surface chemistry and electrical conductivity.14 Generally, heteroatoms (such as boron,15 nitrogen,16 sulfur17,18) doped into a carbon substrate can effectively modify its surface chemistry and electron structures. Among them, nitrogen doping could remarkably improve the wettability, basic property, adsorptive ability, surface polar and conductivity of carbon materials,19 which has been proved by both theoretical and experimental studies. Sun et al. firstly used nitrogen-doped mesoporous carbon (NC) and sulfur to prepare an NC/S composite cathode.20 Comparing with an AC/S composite, the battery based on the NC/S composite exhibited a higher discharge potential and an initial capacity of 1420 mA h g−1 at a current density of 84 mA g−1 (C/20). However, the nitrogen doping was achieved by high temperature NH3 treatment, with only a low sulfur content of 24%. Xu et al. synthesized porous nitrogen-doped carbon nanotubes from a tubular polypyrrole.21 The initial discharge capacity was 1341 mA h g−1 at a current rate of 1 C and a high reversible capacity of 933 mA h g−1 was retained, even after 50 cycles at the same rate. Nitrogen doped carbon can perform not only as a superior cathode for Li–S batteries, but also as an electrode material for supercapacitors that provides great pseudo capacitance and exhibits excellent performance with the help of the nitrogen functionalities.18 However, among current methods of nitrogen doping, it is common to introduce nitrogen onto carbon frameworks by a post process using ammonia or melamine;18,20,22–24 as mentioned above, these methods result in a dreary process and instable properties given by nitrogen because of its uncontrollable amount and inhomogeneity.
We report herein a one step process to synthesize in situ nitrogen doped mesoporous carbon (NMC) with high specific surface area and large pore volume for high capacity lithium–sulfur batteries. Industrial melamine–formaldehyde resin (MF) and magnesium citrate were employed in the one step carbonization without any activation processes or subsequent extra reactions to introduce nitrogen. The prepared NMC serves as a good electron conductive framework and sulfur stabilizer. Besides, the facile approach to introduce nitrogen onto the carbon framework can guarantee a stable nitrogen content and pore structure under the flexible control of the harsh working conditions. Further, the low-cost MF and magnesium citrate can be largely synthesized, facilitating the practical application of Li–S batteries. In comparison with free doped mesoporous carbon (MC), the NMC exhibits a higher capacity retention. The improvement of electrochemical properties of Li–S batteries could be attributed to the interaction between the nitrogen functionalities on the surface of NMC and polysulfide as well as the enhanced electronic conductivity of the carbon matrix. The advantages of the nitrogen doped mesoporous carbon were investigated in detail.
The galvanostatic charge/discharge tests were carried out in potential range of 1.7–2.8 V with a LAND CT-2001A instrument (Wuhan, China). All capacity values were calculated on the basis of sulfur mass. Cyclic voltammetry (CV) tests and electrochemical impedance spectroscopy (EIS) measurements were conducted with a CHI 660D electrochemical measurement system. The CV tests were performed at a scan rate of 0.1 mV s−1 in the voltage range of 1–3 V. The EIS were measured in the frequency range from 100 kHz to 0.1 Hz at open-circuit voltage. All the electrochemical tests were conducted at room temperature.
To characterize the surface properties of NMC, XPS measurements were performed and the results are exhibited in Fig. 1. The peaks centered at about 285.0, 400.0 and 531.0 eV in all survey spectra correspond to the C1s, N1s and O1s, respectively, thus excluding the presence of any other impurities. The high resolution N1s peaks (Fig. 1b) can be divided into three peaks with binding energy of 398.5 ± 0.3, 400.5 ± 0.3 and 401.6 ± 0.3 eV that correspond to pyridinic N, pyrrolic N and graphitic N, respectively. Further, a schematic model of these N species in carbon frame work is shown in Fig. 1c. The pyridinic N is located at the edges of the carbon layer by substituting a carbon atom on the C6 ring,22 which contributes one pair of lone electrons and induces basic properties to the carbon surface.28 In this case, because of the strong inter-atomic interaction, the NMCs bind favourably with S and polysulfides species, hence further limiting the extent of polysulfides dissolution.28 As for pyrrolic N, the nitrogen atom on a five-membered ring contributes two p-electrons to the π system, which may also improve the adsorption ability of the NMCs. What's more, the graphitic N locates inside the graphitic carbon plane by substituting the carbon atom, bonding with three sp2 carbon atoms, which are regarded as electron acceptors due to the higher electronegativity of N (3.04) compared with C (2.55).29 The enhanced electronegativity will introduce more electrons to the graphitic layers, promoting the electronic conductivity of the carbon.30 As discussed above, introducing nitrogen into the carbon framework enhances the material's electronic conductivity, which can be confirmed through measuring the powder electrical resistivity of the NMC and MC under the same pressure of 100 kg. As shown in Fig. 2, the electronic resistivity decreases from 1.60 Ω to 0.67 Ω cm by nitrogen doping.
![]() | ||
Fig. 1 (a) XPS survey. (b) High-resolution N1s XPS spectrum of NMC. (c) Schematic model of three types of NMC: pyridinic, pyrrolic, and graphitic. |
The porous structure of NMC was analyzed by nitrogen sorption and transmission electron microscopy (TEM). The nitrogen adsorption isotherms of NMC and NMC@S are shown in Fig. 3, and the corresponding textural parameters are summarized in Table 1. According to the IUPAC classification, the sample shows type IV adsorption isotherms with a steep condensation step within a narrow relative pressure range and type H1 hysteresis loops for large cylindrical mesoporous.31 The BET surface area of NMC reaches 1599 m2 g−1, and the BJH pore volume is as high as 1.53 m3 g−1. The BJH pore size distribution shows that NMC possesses pore size centered at 3–4 nm, templated by the MgO from the magnesium citrate precursor. The high surface area and large pore volume of NMC guarantee to accommodate a high loading of nano-scaled sulfur and absorb polysufides strongly, implying an excellent electrochemical performance. As shown in Table 1 and Fig. 3, after the impregnation of sulfur, the pore volume and BET surface area of NMC remarkably reduced to 0.0329 m3 g−1 and 11.25 m2 g−1, respectively, indicating that the sulfur is impregnated into the small pores of the NMC. Fig. 4a and b show TEM images of NMC, and it can be seen that NMC has a disordered mesoporous structure, which can accommodate S and polysufides, while the thin walls of the pores not only separate the material into compartments, but also connect the pores with each other, thus constructing the entire conductive network. Moreover, the walls of the pores act as conductors of electrons and Li+, and the disordered structure shortens the transmission distance of Li+.
![]() | ||
Fig. 3 (a) Nitrogen adsorption/desorption isotherms of NMC at 77 K. (b) BJH pore size distributions of the NMCs before and after sulfur loading. |
The sulfur content of NMC@S was determined by TG analysis. As shown in Fig. 5, the sulfur content is as high as 65%, and the high sulfur content in the composites will ensure a high overall energy density per gram of cathode in Li–S batteries. In addition, differential thermal gravity analysis (DTG) of NMC@S and pristine S were conducted and the results are displayed in Fig. 5. Compared with pristine S, the S in the NMC@S evaporates at a significantly elevated temperature and depressed rate, indicating that the nitrogen doped carbon substrate has a stronger interaction with S, which may enhance the performance of Li–S batteries.
XRD analysis was carried out to further investigate the configuration of S in the NMC@S; the patterns of NMC, NMC@S and pristine S are shown in Fig. 6. Similar to the NMC pattern, there are no obvious reflection peaks of bulk sulfur in the NMC@S curves, which is consistent with amorphous carbon materials, demonstrating that nano-sized S was anchored into the pores by adsorption as a result of the disruption of the S crystal structure.
In order to demonstrate the dispersed state of sulfur after sulfur loading, SEM was conducted on the NMC and NMC@S. By comparison of the SEM images before and after sulfur impregnation, no obvious agglomerations of S particles are detected on the surface of NMC@S. To further confirm the distribution of sulfur within the carbon matrix, elemental X-ray mapping was performed at the same time. In Fig. 4, the sulfur mapping shows a homogeneous distribution of sulfur within NMC@S, which is in conformity with the XRD pattern of NMC@S with no observable reflection peaks of sulfur bulk. These observations further confirmed that sulfur diffuses into the pores of the carbon during the melt-diffusion process, which is consistent with the results of BET and pore size distribution test.
![]() | ||
Fig. 7 (a) Cyclic voltammetry plots at a scan rate of 0.1 mV s−1 for the initial three cycles of NMC@S. (b) EIS before cycling of NMC@S and MC@S. |
Fig. 8 displays the discharge profiles of the NMC@S and MC@S cathode at a current density of 100 mA g−1. Typically, two well-defined discharge potential plateaus, corresponding to the formation of higher-order polysulfides (Li2Sn, 4 ≤ n < 8) at 2.3 V, and lower-order Li2S2/Li2S at 2.0 V, are obviously observed for all samples in the discharge process, which is consistent with two cathodic peaks in the CV curves. The initial discharge capacities of NMC@S and MC@S were 1012 and 898 mA h g−1, respectively, corresponding to 61% and 53% utilization of sulfur, respectively. The relatively high capacity in the NMC@S cathode benefits from the enhanced electronic conductivity after nitrogen doping. The prolonged plateau around 2.0 V contributes to the main discharge capacity, and the potential plateaus of the NMC@S cathode are still stable and distinct after 50 cycles, which indicate the excellent electrochemical performance of NMC@S composites. The diffusion of polysulfides should be attributed to the synergistic effect of the developed porous nanostructure, large specific area and the modified surface chemistry with nitrogen doping.
Fig. 9 shows the cycling performance of Li–S batteries based on the NMC@S and MC@S composite electrodes under a current density of 100 mA g−1. Compared to the MC@S composite electrode, the NMC@S composite electrode delivers a higher discharge capacity and has slower capacity fading. The enhanced cycling performance should be attributed to the introduction of the nitrogen onto mesoporous carbon, which improves the adsorption ability towards polysulfides and retards diffusion of polysufides away from the cathode. Owing to the strong inter-atomic interaction, the NMCs bind favourably with S and polysulfides species, hence further limiting the extent of polysulfides dissolution. In addition, the enhanced cycling performance also profits from the promoted electronic conductivity resulted from nitrogen being doped into the carbon, which is further demonstrated by the lower charge transfer resistance observed in the EIS of the NMC@S electrode compared to that of the MC@S electrode (Fig. 7b). The semicircle diameters in the high frequency region of the Nyquist plots reveal that the charge-transfer kinetics are improved by nitrogen doping, indicating the enhanced electronic conductivity. The results are also in good agreement with the results of the powder electronic resistivity (Fig. 3). Additionally, the reactions in Li–S batteries are complex and the mechanism is still under study;34 the cycling stability requires improvement in our subsequent study.
This journal is © The Royal Society of Chemistry 2015 |